21.12.2013 Views

Conversion of pharmaceuticals and other drugs by fungal ... - IHI Zittau

Conversion of pharmaceuticals and other drugs by fungal ... - IHI Zittau

Conversion of pharmaceuticals and other drugs by fungal ... - IHI Zittau

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Conversion</strong> <strong>of</strong> <strong>pharmaceuticals</strong> <strong>and</strong> <strong>other</strong> <strong>drugs</strong> <strong>by</strong><br />

<strong>fungal</strong> peroxygenases<br />

Umsetzung von Pharmazeutika und psychoaktiven<br />

Substanzen mit pilzlichen Peroxygenasen<br />

Vom Institutsrat des Internationalen Hochschulinstitutes <strong>Zittau</strong> und der<br />

Fakultät Mathematik und Naturwissenschaften der Technischen Universität<br />

Dresden<br />

genehmigte<br />

DISSERTATION<br />

Zur Erlangung des akademischen Grades<br />

Doctor rerum naturalium<br />

(Dr.rer.nat.)<br />

von<br />

geboren am<br />

vorgelegt<br />

Marzena Poraj-Kobielska, Dipl.-Ing. (Umwelttechnik)<br />

25 Juli 1982 in Katowice (Polen)<br />

Gutachter:<br />

Herr Pr<strong>of</strong>. Dr. Martin H<strong>of</strong>richter (TU Dresden-<strong>IHI</strong> <strong>Zittau</strong>)<br />

Herr Pr<strong>of</strong>. Dr. Kenneth E. Hammel (University <strong>of</strong> Madison)<br />

Herr Pr<strong>of</strong>. Dr.-Ing. Korneliusz Miksch (Politechnika Śląska,<br />

Gliwice)<br />

Tag der Verteidigung: 26 April 2013


<strong>Conversion</strong> <strong>of</strong> <strong>pharmaceuticals</strong> <strong>and</strong> <strong>other</strong> <strong>drugs</strong> <strong>by</strong><br />

<strong>fungal</strong> peroxygenases<br />

Approved <strong>by</strong> the Council <strong>of</strong> International Institute <strong>Zittau</strong> <strong>and</strong> the Faculty <strong>of</strong><br />

Science <strong>of</strong> the TU Dresden<br />

Academic Dissertation<br />

Doctor rerum naturalium<br />

(Dr.rer.nat.)<br />

<strong>by</strong><br />

born on<br />

Marzena Poraj-Kobielska, Dipl.-Ing. (Environmental Engineering)<br />

July 25, 1982 in Katowice (Pol<strong>and</strong>)<br />

Reviewers: Pr<strong>of</strong>. Dr. Martin H<strong>of</strong>richter (TU Dresden-<strong>IHI</strong> <strong>Zittau</strong>)<br />

Pr<strong>of</strong>. Dr. Kenneth E. Hammel (University <strong>of</strong> Madison)<br />

Pr<strong>of</strong>. Dr.-Ing. Korneliusz Miksch (Politechnika Śląska, Gliwice)<br />

Day <strong>of</strong> defence: April 26, 2013


CONTENTS<br />

Contents<br />

Contents…………………………………………………………………………………….I<br />

List <strong>of</strong> Abbreviations…………………………………………………………………….IV<br />

Zusammenfassung……………………………………………………………………….VI<br />

Summary……………………………………………………………………………..…VIII<br />

Streszczenie………………………………………………………………………………..X<br />

1 Introduction ................................................................................................................... 1<br />

1.1 Physiological <strong>and</strong> environmental aspects <strong>of</strong> <strong>pharmaceuticals</strong>............................... 1<br />

1.1.1 Pharmaceuticals............................................................................................. 1<br />

1.1.2 Pharmacokinetics <strong>and</strong> Metabolism................................................................ 2<br />

1.1.3 Pharmaceutical <strong>drugs</strong> development <strong>and</strong> studies ............................................ 5<br />

1.1.4 Environmental aspects................................................................................... 6<br />

1.2 Cytochrome P450 enzymes ................................................................................. 10<br />

1.2.1 Classification <strong>of</strong> electron transport systems <strong>of</strong> P450s................................. 13<br />

1.2.2 Bacterial P450s............................................................................................ 15<br />

1.2.3 Plant P450s .................................................................................................. 16<br />

1.2.4 Fungal P450s ............................................................................................... 17<br />

1.2.5 Animal P450s .............................................................................................. 20<br />

1.2.6 Engineered P450s ........................................................................................ 21<br />

1.2.7 Catalyzed reactions...................................................................................... 22<br />

1.2.8 Reaction mechanism <strong>of</strong> P450s..................................................................... 23<br />

1.3 Peroxygenases ..................................................................................................... 27<br />

1.3.1 Fungal unspecific peroxygenases ................................................................ 30<br />

1.4 Aims <strong>and</strong> objectives............................................................................................. 31<br />

2 Materials <strong>and</strong> Methods ................................................................................................ 33<br />

2.1 Chemicals <strong>and</strong> reactants ...................................................................................... 33<br />

2.1.1 Pharmaceuticals <strong>and</strong> their metabolites ........................................................ 33<br />

2.1.2 Drugs <strong>and</strong> their metabolites......................................................................... 33<br />

2.1.3 Steroids <strong>and</strong> anabolic agents........................................................................ 34<br />

2.1.4 Benzodioxole derivatives ............................................................................ 34<br />

2.2 Fungal cultivation <strong>and</strong> purification <strong>of</strong> peroxygenases......................................... 34<br />

2.2.1 Agrocybe aegerita <strong>and</strong> its peroxygenase..................................................... 34<br />

2.2.2 Coprinellus radians <strong>and</strong> its peroxygenase................................................... 38<br />

2.2.3 Marasmius rotula <strong>and</strong> its peroxygenase...................................................... 41<br />

2.3 UV-Vis spectroscopy........................................................................................... 42<br />

2.3.1 Veratryl alcohol assay for peroxygenases ................................................... 42<br />

2.3.2 ABTS assay for laccase ............................................................................... 43<br />

2.3.3 Nitrobenzodioxole assay for peroxygenases ............................................... 43<br />

2.3.4 Determination <strong>of</strong> protein concentration....................................................... 45<br />

2.3.5 Difference spectra........................................................................................ 46<br />

2.4 St<strong>and</strong>ard reaction conditions for peroxygenase conversions............................... 46<br />

2.5 Product identification .......................................................................................... 47<br />

2.5.1 HPLC-Method I........................................................................................... 47<br />

2.5.2 UHPLC-Method .......................................................................................... 47<br />

2.5.3 HPLC-Method for chiral separation............................................................ 47<br />

2.5.4 HPLC-MS Method I .................................................................................... 47<br />

2.5.5 HPLC-MS Method II................................................................................... 48<br />

2.5.6 HPLC-MS Method III ................................................................................. 48<br />

2.5.7 Mass spectrometry for HPLC-MS methods ................................................ 48<br />

I


2.5.8 IC-Method for organic acids ....................................................................... 48<br />

2.6 Stoichiometrical analyses.................................................................................... 48<br />

2.6.1 Sildenafil ..................................................................................................... 48<br />

2.6.2 Safrole ......................................................................................................... 49<br />

2.7 Enzyme kinetics .................................................................................................. 49<br />

2.7.1 Pharmaceuticals........................................................................................... 49<br />

2.7.2 Nitrobenzodioxole....................................................................................... 49<br />

2.8 Experiments with 18 O-isotopes............................................................................ 50<br />

2.8.1 Pharmaceuticals........................................................................................... 50<br />

2.8.2 Nitrobenzodioxole....................................................................................... 50<br />

2.9 Deuterium isotope effect experiment .................................................................. 50<br />

2.10 Experiment with different peroxides <strong>and</strong> peroxide sources ................................ 51<br />

2.11 In-vivo incubation experiment............................................................................. 51<br />

2.12 <strong>Conversion</strong> <strong>of</strong> propranolol at environmentally relevant concentration............... 53<br />

2.13 Long term experiments........................................................................................ 53<br />

3 Results ......................................................................................................................... 55<br />

3.1 Aromatic ring hydroxylation............................................................................... 55<br />

3.1.1 Ring hydroxylation at <strong>pharmaceuticals</strong>....................................................... 55<br />

3.1.2 Ring hydroxylation at <strong>drugs</strong> <strong>of</strong> abuse.......................................................... 58<br />

3.2 Oxidation <strong>of</strong> alkyl side chains............................................................................. 59<br />

3.3 O-Dealkylation .................................................................................................... 62<br />

3.3.2 Ester cleavage.............................................................................................. 64<br />

3.4 N-dealkylation ..................................................................................................... 64<br />

3.5 O-Demethylenation ............................................................................................. 68<br />

3.6 Oxidation <strong>of</strong> steroids ........................................................................................... 69<br />

3.7 Stoichiometrical analyses.................................................................................... 71<br />

3.7.1 Sildenafil ..................................................................................................... 71<br />

3.7.2 Safrole ......................................................................................................... 72<br />

3.8<br />

18 O-Isotope incorporation.................................................................................... 72<br />

3.9 Kinetics experiments ........................................................................................... 73<br />

3.9.1 Michaelis-Menten kinetics .......................................................................... 73<br />

3.9.2 Bisubstrate kinetics...................................................................................... 74<br />

3.10 Deuterium isotope effect experiments................................................................. 75<br />

3.11 Spectrophotometric nitrobenzodioxole assay...................................................... 75<br />

3.12 <strong>Conversion</strong> <strong>of</strong> a set <strong>of</strong> human P450 substrates.................................................... 78<br />

3.13 Difference spectroscopy...................................................................................... 81<br />

3.14 Different peroxides <strong>and</strong> peroxide-generating systems ........................................ 82<br />

3.15 In-vivo experiments ............................................................................................. 83<br />

3.16 Tests with environmentally relevant concentrations <strong>of</strong> propranolol ................... 85<br />

3.17 Long-term experiments ....................................................................................... 86<br />

3.17.1 Pre-study with Vivaspin® (fed-batch method) ........................................... 86<br />

3.17.2 Long-term experiment using a hollow fiber cartridge................................. 87<br />

3.18 Scope <strong>of</strong> peroxygenase reactions concerning <strong>pharmaceuticals</strong> <strong>and</strong> <strong>drugs</strong>.......... 88<br />

4 Discussion ................................................................................................................... 90<br />

4.1 Preparation <strong>of</strong> human drug metabolites............................................................... 90<br />

4.2 Mechanistic studies ............................................................................................. 99<br />

4.2.1 Source <strong>of</strong> oxygen <strong>and</strong> stoichiometry........................................................... 99<br />

4.2.2 Deuterium isotope effect ........................................................................... 100<br />

4.2.3 Kinetics...................................................................................................... 100<br />

4.2.4 Assays for human cytochrome P450 activities.......................................... 102<br />

4.2.5 Peroxides <strong>and</strong> peroxides generation systems ............................................ 104<br />

4.2.6 Difference binding spectra ........................................................................ 106<br />

4.2.7 Phenoxyl radical scavenger....................................................................... 106<br />

4.2.8 Stoichiometry <strong>and</strong> proposed mechanism <strong>of</strong> demethylenation................... 107<br />

II


CONTENTS<br />

4.3 NBD assay ......................................................................................................... 109<br />

4.4 Long term <strong>and</strong> in-vivo experiments................................................................... 110<br />

4.5 Homology model structures <strong>of</strong> APOs................................................................ 113<br />

4.6 Key findings ...................................................................................................... 120<br />

4.7 Outlook .............................................................................................................. 123<br />

5 References ................................................................................................................. 125<br />

6 Appendix ................................................................................................................... 143<br />

6.1 Aromatic ring hydroxylation ............................................................................. 143<br />

6.2 Oxidation <strong>of</strong> alkyl side chains ........................................................................... 144<br />

6.3 O-Dealkylation .................................................................................................. 144<br />

6.4 N-dealkylation ................................................................................................... 145<br />

6.5 Oxidation <strong>of</strong> steroids ......................................................................................... 145<br />

6.6 O-Demethylenation ........................................................................................... 146<br />

6.7 Spectrophotometric nitrobenzodioxole assay.................................................... 146<br />

6.8 Stoichiometrical analyses .................................................................................. 147<br />

6.9 Experiment with propranolol in environmentally relevant concentrations ....... 148<br />

6.10 List <strong>of</strong> publications ............................................................................................ 150<br />

III


LIST OF ABBREVIATIONS<br />

List <strong>of</strong> Abbreviations<br />

Aae Agrocybe aegerita<br />

Ala Alanine<br />

APOs aromatic/unspecific peroxygenases<br />

Arg Arginine<br />

BPs By-products<br />

Cfu Caldariomyces (Leptoxyphium) fumago<br />

CPO Chloroperoxidase<br />

Cra Coprinellus radians<br />

Cve Coprinopsis verticillata<br />

DAD Diod array detector<br />

DNPH 2,4-Dinitrophenylhydrazine (Brady's reagent)<br />

EC Enzyme Commission<br />

ESI Electrospray Ionization<br />

FPLC Fast Protein Liquid Chromatography<br />

Glu Glutamic acid<br />

Gly Glycine<br />

HDMs Human drug metabolites<br />

HPLC High Performance Liquid Chromatography<br />

IC<br />

Ion exchange Chromatography<br />

IEF Isoelectric Focussing<br />

Ile<br />

Isoleucine<br />

k cat<br />

k cat /K m<br />

kDa<br />

((k H /k D ) obs )<br />

Leu<br />

LRET<br />

mCPBA<br />

MEA<br />

MPO<br />

Mro<br />

IV<br />

turnover number<br />

catalytic efficiency<br />

kilo Dalton<br />

observed deuterium isotope effect<br />

(k H /k D ) intrinsic deuterium isotope effect<br />

K m<br />

Michaelis-Menten-(Henri) constant<br />

Leucine<br />

Long-Range Electron Transfer<br />

meta-Chloroperbenzoic acid<br />

Malt-extract agar<br />

Myeloperoxidase<br />

Marasmius rotula


LIST OF ABBREVIATIONS<br />

MS<br />

NADH<br />

NADPH<br />

NBD<br />

NSAIDs<br />

P450 BM3<br />

P450 BSβ<br />

P450 cam<br />

P450 foxy<br />

P450 lun<br />

P450 nor<br />

P450s<br />

P450 SPα<br />

Phe<br />

pI<br />

PPs<br />

Pro<br />

SDS-PAGE<br />

SEC<br />

Ser<br />

Thr<br />

Tyr<br />

UHPLC<br />

UV<br />

VA<br />

Val<br />

Vis<br />

WWTPs<br />

Mass Spectrometry<br />

Nicotinamide Adenine Dinucleotide<br />

Nicotinamide Adenine Dinucleotide Phosphate<br />

Nitrobenzodioxole<br />

Non-steroidal anti-inflammatory <strong>drugs</strong><br />

P450-reductase fusion enzyme from Bacillus megaterium<br />

Fatty acid peroxygenase from Bacillus subtilis<br />

Camphor 5-Monooxygenase from Pseudomonas<br />

P450 from Fusarium oxysporum<br />

P450 from Curvularia lunata<br />

Nitric oxide reductase<br />

Cytochrome P450 Monooxygenases<br />

Fatty acid peroxygenase from Sphinomonas paucinobilis<br />

Phenylalanine<br />

Isoelectric point<br />

Pharmaceutical products<br />

Proline<br />

Sodium dodecyl sulfate polyacrylamide<br />

Size Exclusion Chromatography<br />

Serine<br />

Threonine<br />

Tyrosine<br />

Ultra High Performance Liquid Chromatography<br />

Ultraviolet<br />

Veratryl alcohol<br />

Valine<br />

Visible<br />

Waste Water Treatment Plants<br />

V


SUMMARY<br />

Zusammenfassung<br />

In den letzten Jahren sind Pharmazeutika und deren Metabolite mehr und mehr in den<br />

Fokus der Wissenschaft gerückt. Diese Substanzen sind aufgrund ihrer physiologischen<br />

und toxikologischen sowie ökotoxikologischen Wirkungen im menschlichen Körper bzw.<br />

in der Umwelt von besonderem Interesse. Cytochrom-P450-Enzyme (P450s) spielen eine<br />

Schlüsselrolle bei der Umsetzung und Detoxifizierung bioaktiver Substanzen, darunter<br />

vieler Pharmazeutika und Drogen. Es h<strong>and</strong>elt sich bei diesen Enzymen in erster Linie um<br />

Monooxygenasen, die intrazellulär lokalisiert und relativ instabil sind; sie benötigen<br />

komplexe, teure K<strong>of</strong>aktoren und sind nur unter hohem Aufw<strong>and</strong> zu reinigen, was ihre<br />

Anwendung in isolierter Form insgesamt erschwert. Die hier durchgeführten<br />

Untersuchungen zu pilzlichen Peroxygenasen haben gezeigt, dass diese Enzymsubklasse<br />

(EC 1.11.2.x) ein hohes Oxyfunktionalisierungspotenzial besitzt und eine Vielzahl P450-<br />

typischer Reaktionen zu katalysieren vermag. Peroxygenasen sind extrazelluläre, d.h.<br />

sekretierte Pilzenzyme, die eine hohe Stabilität aufweisen und lediglich ein Peroxid als<br />

Kosubstrat benötigen. Die unter Verwendung der unspezifischen/aromatischen<br />

Peroxygenasen (APOs) von Agrocybe aegerita, Coprinellus radians und Marasmius rotula<br />

gewonnenen Ergebnisse belegen, dass APOs verschiedene H 2 O 2 -abhängige<br />

Monooxygenierungen von Pharmazeutika und psychoaktiven Substanzen realisieren. Dazu<br />

gehören i) die Monooxygenierung von Aromaten, ii) die benzylische Hydroxylierung von<br />

Toluolderivaten, iii) die O-Dealkylierung verschiedener Etherstrukturen einschließlich der<br />

Spaltung von Benzodioxolen (O-Demethylenierung) und Estern sowie iv) die N-<br />

Dealkylierung von sekundären und tertiären Aminen. Die untersuchten Peroxygenasen<br />

wiesen teilweise deutliche Unterschiede im Substratspektrum und den präferierten<br />

Oxidationspositionen auf. Dieser Befund eröffnet die Möglichkeit, zukünftig einen<br />

„enzymatischen Werkzeugkasten“ auf Basis pilzlicher Peroxygenasen für die<br />

Oxyfunktionalisierung von pharmazeutisch relevanten Wirkst<strong>of</strong>fen zu entwickeln.<br />

Mechanistische Experimente zeigten, dass (1) die Monooxygenierungen stets unter Einbau<br />

eines aus dem Peroxid stammenden Sauerst<strong>of</strong>fatoms erfolgen, (2) die Deethylierung von<br />

Phenacetin-d1 einen Deuteriumisotopeneffekt ähnlich dem der P450s aufweist, (3) die<br />

katalytischen Effizienzen für die untersuchten Oxidationen im gleichen Bereich wie die der<br />

P450s liegen (wobei die k cat - und K m -Werte deutlich höher ausfallen), (4) die kinetischen<br />

Untersuchungen zur Oxidation von Nitro-1,3-Benzodioxol parallele Verläufe der<br />

ermittelten Ausgleichsgeraden in der doppelt reziproken Darstellung ergaben, was für<br />

einen “Ping-Pong-Mechanismus“ spricht, (5) sich das Substratspektrum und die<br />

Aktivitätsmuster der APOs in einem weiten Bereich mit denen der wichtigsten<br />

VI


SUMMARY<br />

menschlichen P450s decken sowie dass (6) die in Bindungsstudien gewonnenen<br />

Differenzspektren denen des Phenoltyps der P450s entsprechen. Desweiteren erwiesen sich<br />

APOs in Langzeitexperimenten über zwei Wochen als stabil und aktiv und sie waren in der<br />

Lage, Pharmazeutika in umweltrelevanten Konzentrationen (ppb-Bereich) zu oxidieren.<br />

All die genannten Eigenschaften legen nahe, dass APOs eine interessante Alternative zur<br />

enzymatischen Umsetzung von Pharmazeutika sowie zur Herstellung von humanen<br />

Pharmazeutika-Metaboliten darstellen, die z.B. Einsatz in der medizinischpharmakologischen<br />

Forschung oder im Umweltbereich (Entfernung von Pharmazeutika<br />

aus Umweltmedien) finden könnten.<br />

VII


SUMMARY<br />

Summary<br />

Over the recent years, increasing scientific attention has been paid to <strong>pharmaceuticals</strong>,<br />

<strong>other</strong> <strong>drugs</strong> <strong>and</strong> their metabolites. These substances are <strong>of</strong> particular interest because <strong>of</strong><br />

their physiological, toxicological <strong>and</strong> ecotoxicological effects in the human body <strong>and</strong><br />

respectively in the environment. Cytochrome P450 enzymes (P450s) play a key role in the<br />

conversion <strong>and</strong> detoxification <strong>of</strong> bioactive compounds including many <strong>pharmaceuticals</strong><br />

<strong>and</strong> <strong>drugs</strong>. Most <strong>of</strong> these enzymes belong to the monooxygenases; they are intracellular<br />

<strong>and</strong> rather unstable biocatalysts that are difficult to purify <strong>and</strong> require expensive, complex<br />

c<strong>of</strong>actors, which alltogether hampers their use in isolated form. The investigations carried<br />

out here with <strong>fungal</strong> peroxygenases have shown that this enzyme sub-subclass (EC<br />

1.11.2.x) has a promising potential for oxyfunctionalizations <strong>and</strong> can catalyze a variety <strong>of</strong><br />

reactions typical for P450s. Peroxygenases are extracellular, i.e. secreted <strong>fungal</strong> enzymes<br />

with high stability, which merely need peroxide for function. Results obtained with the<br />

unspecific/aromatic peroxygenases (APOs) <strong>of</strong> Agrocybe aegerita, Coprinellus radians <strong>and</strong><br />

Marasmius rotula have demonstrated that APOs catalyze numerous H 2 O 2 -dependent<br />

monooxygenations <strong>of</strong> <strong>pharmaceuticals</strong> <strong>and</strong> psychoactive <strong>drugs</strong>. Among them are i) the<br />

monooxygenation <strong>of</strong> aromatic compounds, ii) the benzylic hydroxylation <strong>of</strong> toluene<br />

derivatives, iii) the O-dealkylation <strong>of</strong> different ether structures including the scission <strong>of</strong><br />

benzodioxoles (O-demethylenation) <strong>and</strong> esters as well as iv) the N-dealkylation <strong>of</strong><br />

secondary <strong>and</strong> tertiary amines. The peroxygenases studied considerably differ in their<br />

substrate spectrum <strong>and</strong> the preferred positions <strong>of</strong> oxidation. This finding opens the<br />

possibility to develop in the future an “enzymatic toolbox“ on the basis <strong>of</strong> <strong>fungal</strong><br />

peroxygenases for the oxyfunctionalization <strong>of</strong> pharmaceutically relevant compounds.<br />

Mechanistic studies showed that (1) the monooxygenations always proceed via<br />

incorporation <strong>of</strong> one oxygen atom from the peroxide, (2) the demethylation <strong>of</strong> phenacetind1<br />

established a deuterium isotope effect similar to P450s, (3) the catalytic efficiencies for<br />

the studied oxidations are in the same range as those <strong>of</strong> P450s (though the k cat - <strong>and</strong> K m -<br />

values are noticeably higher), (4) the kinetic studies with nitro-1,3-benzodioxole gave<br />

parallel double reciprocal plots suggestive <strong>of</strong> a “ping pong” mechanism, (5) the substrate<br />

spectrum <strong>and</strong> the activity pattern <strong>of</strong> APOs follows in a wide range those <strong>of</strong> the human key<br />

P450s as well as that (6) the difference spectra obtained in bindings studies are <strong>of</strong> the<br />

phenol type <strong>of</strong> P450s. Furthermore, APOs were found to be stable <strong>and</strong> active in long term<br />

experiments over two weeks <strong>and</strong> they oxidized <strong>pharmaceuticals</strong> at low, environmentally<br />

relevant concentration (ppb range). All the above properties strongly indicate that APOs<br />

respresent an interesting alternative for the enzymatic conversion <strong>of</strong> <strong>pharmaceuticals</strong> as<br />

VIII


SUMMARY<br />

well as for the preparation <strong>of</strong> human drug metabolites, for example, in medicinal <strong>and</strong><br />

pharmacological research or the bioremediation sector (removal <strong>of</strong> <strong>pharmaceuticals</strong> from<br />

environmental media).<br />

IX


SUMMARY<br />

Streszczenie<br />

W ostatnich latach, w świecie nauki, coraz więcej uwagi poświęca się lekom i ich<br />

metabolitom. Substancje te cieszą się szczególnym zainteresowaniem z powodu ich<br />

fizjologicznego oraz toksykologicznego działania w ludzkim organiźmie, a także na<br />

środowisko naturalne. Cytochromy P450 (P450s) odgrywają kluczową rolę w konwersji i<br />

detoksyfikacji bioaktywnych substancji, między innymi wielu leków i substancji<br />

psychoaktywnych. W pierwszym rzędzie należą do tej grupy enzymów monooksygenazy,<br />

które są wewnątrzkomórkowe i relatywnie niestabilne. Potrzebują kompleksowych,<br />

drogich k<strong>of</strong>aktorów i ich oczyszczanie wymaga dużego nakładu pracy, co utrudnia ich<br />

zastosowanie w wyizolowanej formie. Przeprowadzone badania z grzybowymi<br />

peroksygenazami pokazały, że ta podklasa enzymów (EC 1.11.2.x) posiada wysoki<br />

potencjał ox<strong>of</strong>unkcjonalizacyjny i katalizuje wiele typowych reakcji dla cytochromów<br />

P450. Peroksygenazy są zewnątrzkomórkowe, to znaczy należą do wydzielanych<br />

enzymów grzybowych. Charakteryzują się wysoką stabilnością i wymagają jedynie<br />

obecności nadtlenku wodoru jako kosubstratu. Wyniki otrzymane przy użyciu<br />

aromatycznych/niespecyficznych peroksygenaz (APOs) z Agrocybe aegerita, Coprinellus<br />

radians i Marasmius rotula pokazują, że APOs katalizują wiele rożnych, zależnych od<br />

H 2 O 2 , monooksygenacji leków i substancji psychoaktywnych. Do tego należą i)<br />

monooksygenacja aromatów, ii) hydroksylacja grupy benzylowej pochodnych fenolu, iii)<br />

O-dealkilacja różnych struktur eterycznych włączając rozszczepianie benzodioksoli (Odemetylenowanie)<br />

oraz estrów, jak rownież iv) N-dealkilacja. Badane peroksygenazy<br />

wykazują częściowo znaczne różnice w spektrum substratów i preferowanej pozycji<br />

oksydacji. Powyższe wyniki otwierają możliwość opracowania w przyszłości tak zwanej<br />

„enzymatycznej skrzynki narzędziowej“ na bazie grzybowych peroksygenaz do<br />

ox<strong>of</strong>unkcjonalizowania farmaceutycznie istotnych substancji czynnych. Mechanistyczne<br />

eksperymenty pokazały, że (1) monooksygenacje nastepują przez wbudowanie atomu tlenu<br />

pochodzącego z nadtlenku wodoru, (2) demetylowanie fenacetyny-d1 wykazuje podobny<br />

efekt izotopowy deuteru jak cytochromy P450, (3) wydajność katalityczna dla badanych<br />

oksydacji znajduje się w podobnym zakresie jak dla P450s, przy czym wartości k cat i K m są<br />

znacznie wyższe, (4) kinetyczne badania w zakresie oksydacji nitro-1,3-benzodioxolu<br />

wykazały równoległe przebiegi prostych w podwójnym wykresie Lineweavera-Burka, co<br />

odpowiada tak zwanemu mechanizmowi “Ping-Pong“, (5) spektrum substratów i rodzaj<br />

aktywności w przypadku APOs pokrywa się w dużym zakresie z najważniejszmi ludzkimi<br />

P450s, oraz że (6) badania w zakresie wiązań wykazały spektra różnicowe w dużym<br />

stopniu przypominające spektra fenolowe cytochromów P450. Ponadto w eksperymentach<br />

X


SUMMARY<br />

długoterminowych (dwa tygodnie) grzybowe peroksygenazy okazały się <strong>by</strong>ć stabilne<br />

i aktywne. Były również w stanie oksydować leki, w koncentracjach wystepujących<br />

w środowisku naturalnym (stężenie w zakresie ppb).Wszystkie wymienione wyżej<br />

właściwości wskazują na to, że grzybowe peroksygenazy stanowią interesującą<br />

alternatywę do enzymatycznej konwersji leków, a także do produkcji ludzkich<br />

metabolitów. Mogą <strong>by</strong>ć na przykład stosowane w badaniach medyczno-farmakologicznych<br />

oraz w obrębie ochrony środowiska (usuwanie leków ze środowiska).<br />

XI


INTRODUCTION<br />

1 Introduction<br />

1.1 Physiological <strong>and</strong> environmental aspects <strong>of</strong> <strong>pharmaceuticals</strong><br />

1.1.1 Pharmaceuticals<br />

On 31 March 2004, the European Parliament <strong>and</strong> the Council <strong>of</strong> the European Union<br />

defined a medicinal product: "(a) Any substance or combination <strong>of</strong> substances presented as<br />

having properties for treating or preventing disease in human beings; or (b) Any substance<br />

or combination <strong>of</strong> substances which may be used in or administered to human beings either<br />

with a view to restoring, correcting or modifying physiological functions <strong>by</strong> exerting a<br />

pharmacological, immunological or metabolic action, or to making a medical diagnosis<br />

"(Parliment <strong>and</strong> Council 2004). Pharmaceuticals from a wide spectrum <strong>of</strong> therapeutic<br />

classes are used worldwide in human <strong>and</strong> animal medicine. Pharmaceuticals include more<br />

then 4,000 molecules with different physico-chemical <strong>and</strong> biological properties <strong>and</strong><br />

distinct modes <strong>of</strong> biochemical action (Beausse 2004, Mompelat et al. 2009, Whitacre et al.<br />

2010). Pharmaceuticals can be classified <strong>by</strong> chemical properties, mode or route <strong>of</strong><br />

administration, biological system affected or therapeutic effects. An elaborate <strong>and</strong> widely<br />

used classification system is the Anatomical Therapeutic Chemical Classification System<br />

(ATC system) (Fricke et al. 2012, WHO Collaborating Centre for Drug Statistics<br />

Methodology 2012). Important classes <strong>of</strong> <strong>pharmaceuticals</strong> include:<br />

⎯ Analgesics (painkillers),<br />

⎯ Antibiotics (inhibiting microbial growth),<br />

⎯ Antihypertensives/α-/β-blockers (α-/β-adrenoreceptor antagonists; used to treat<br />

hypertension = high blood pressure),<br />

⎯ Antiepileptic <strong>drugs</strong> (enhance the effect <strong>of</strong> the neurotransmitter γ-aminobutyric<br />

acid),<br />

⎯ Nonsteroidal anti-inflammatory <strong>drugs</strong>, NSAIDs (with analgesic <strong>and</strong> antipyretic<br />

(fever-reducing) effects; in higher doses with anti-inflammatory effects),<br />

⎯ Antitussives (cough medicines),<br />

⎯ Bronchodilators/antiasthmatic <strong>drugs</strong> (β2-agonists, anticholinergics, <strong>and</strong><br />

theophylline),<br />

⎯ Psychoactive <strong>drugs</strong> (antidepressants, stimulants, antipsychotics, mood stabilizers,<br />

anxiolytics, depressants, hallucinogens),<br />

⎯ Chem<strong>other</strong>apeutic agents (anticancers), <strong>and</strong><br />

1


INTRODUCTION<br />

⎯ Hormones (contraceptives, anabolics).<br />

The World Health Organization keeps a list <strong>of</strong> essential medicines. Essential medicines are<br />

defined as: "those <strong>drugs</strong> that satisfy the health care needs <strong>of</strong> the majority <strong>of</strong> the population;<br />

they should therefore be available at all times in adequate amounts <strong>and</strong> in appropriate<br />

dosage forms, at a price the community can afford." (World Health Organization 2012).<br />

The core list presents a number <strong>of</strong> minimum medicine needs for a basic health-care system,<br />

listing the most efficacious, safe <strong>and</strong> cost-effective medicines for priority conditions.<br />

Priority conditions are selected on the basis <strong>of</strong> current <strong>and</strong> estimated future public health<br />

relevance, <strong>and</strong> potential for safe <strong>and</strong> cost-effective treatment. The complementary list<br />

presents essential medicines for priority diseases, for which specialized diagnostic or<br />

monitoring facilities, <strong>and</strong>/or specialist medical care, <strong>and</strong>/or specialist training are needed.<br />

In case <strong>of</strong> doubt, medicines may also be listed as complementary on the basis <strong>of</strong> consistent<br />

higher costs or less attractive cost-effectiveness in a variety <strong>of</strong> settings (World Health<br />

Organization 2011). As on 17 June 2011, the BfArM (Bundesamt für Arzneimittel und<br />

Medizinprodukte; the Federal Institute for Drugs <strong>and</strong> Medical Devices) reported marketing<br />

authorizations or registrations for 60,237 pharmaceutical <strong>drugs</strong> in all indications. The<br />

"Rote Liste®" for 2011, the comprehensive drug registry for Germany, includes only 8,280<br />

preparations, with a total <strong>of</strong> 33,737 prices (<strong>pharmaceuticals</strong> are almost always marketed at<br />

different prices for different package sizes). This difference between nearly 60,000<br />

authorized <strong>pharmaceuticals</strong> on the one h<strong>and</strong> <strong>and</strong> only roughly 10,000 drug entries in the<br />

"Rote Liste®" can be explained <strong>by</strong> the different tallying methods <strong>and</strong> the only partial<br />

listing <strong>of</strong> self-medication <strong>drugs</strong> in the "Rote Liste®" (German Association for<br />

Pharmaceutical Industry 2011).<br />

1.1.2 Pharmacokinetics <strong>and</strong> Metabolism<br />

Pharmacokinetics, "the movement <strong>of</strong> <strong>drugs</strong>" is the study <strong>of</strong> a drug <strong>and</strong>/or its metabolite<br />

kinetics in the body. The human body is a very complex system <strong>and</strong> a drug undergoes<br />

many steps as it is being absorbed (A), distributed through the body (D), metabolized (M),<br />

<strong>and</strong>/or excreted (E) (figure 1). When the studies are focused solely on one specific<br />

pharmacokinetic aspect (A,D,M or E) <strong>by</strong> in vitro, in situ, in vivo or in silico techniques<br />

they are usually referred to as ADME studies, whereas the term pharmacokinetics is<br />

normally reserved for in vivo studies where all four aspects are considered (Ruiz-Garcia et<br />

al. 2008). Biotransformation/metabolism is an integral component <strong>of</strong> the processes that<br />

govern the pharmacokinetic pr<strong>of</strong>ile <strong>of</strong> therapeutically used <strong>drugs</strong>. Biotransformation<br />

generally converts nonpolar, lipophilic pharmacologically active drug molecules into polar,<br />

2


INTRODUCTION<br />

inactive or nontoxic metabolites that are readily eliminated <strong>by</strong> kidneys or <strong>other</strong> organs. The<br />

liver is the major site <strong>of</strong> drug metabolism, with intestinal drug extraction playing a<br />

secondary yet potentially important role in the presystemic elimination <strong>of</strong> orally<br />

administered agents. Traditionally, biotransformation pathways are classified into two<br />

groups:<br />

⎯ Phase I reactions that involve (oxy)functionalization <strong>of</strong> drug substrates to yield<br />

more polar derivatives such as alcohols, phenols or carboxylic acids, <strong>and</strong><br />

⎯ Phase II reactions involving bimolecular conjugation with endogenous hydrophilic<br />

moieties to yield products such as glucuronides <strong>and</strong> sulfates that are readily<br />

excreted <strong>by</strong> kidneys (Venkatakrishnan et al. 2001).<br />

Figure 1 Processes involved in the determination <strong>of</strong> pharmacokinetic parameters. Adapted from<br />

Gibson <strong>and</strong> Skett (Gibson <strong>and</strong> Skett 2008).<br />

Cytochrome P450s (P450s) are the enzymes principally responsible for Phase I oxidations<br />

in animals. Because the chemicals that must be metabolized are structurally very diverse,<br />

animals express multiple P450s that have evolved via gene duplications <strong>and</strong> conversions,<br />

thus giving rise to genes encoding new P450 forms that metabolize not only xenobiotics,<br />

but also endogenous substances. One consequence <strong>of</strong> human P450 gene evolution is the<br />

polymorphism <strong>of</strong> drug metabolism, leading to marked differences in the response <strong>of</strong><br />

individuals to the toxic <strong>and</strong> carcinogenic effects <strong>of</strong> <strong>drugs</strong> <strong>and</strong> <strong>other</strong> environmental<br />

3


INTRODUCTION<br />

chemicals. Differences in P450 expression between human individuals, which are<br />

correlated with ethnicity in some cases, may reflect the presence or absence <strong>of</strong> a particular<br />

P450 gene, but may also involve transcriptional regulation. These differences may be<br />

clinically manifested in quantitative differences in the metabolism <strong>of</strong> certain <strong>drugs</strong>,<br />

resulting in "poor" or "extensive" metabolizers (Souček <strong>and</strong> Gut 1992). P450-mediated<br />

drug metabolism can also lead to altered drug efficacies through inactivation <strong>of</strong> an active<br />

drug or activation <strong>of</strong> a prodrug (Ding <strong>and</strong> Kaminsky 2003). Three P450 gene families<br />

(CYP1, CYP2, <strong>and</strong> CYP3) <strong>and</strong> to some degree a fourth one (CYP4) are currently thought<br />

to be responsible for the majority <strong>of</strong> hepatic drug metabolism in humans (Wrighton <strong>and</strong><br />

Stevens 1992).<br />

Figure 2 A summary <strong>of</strong> major human P450s involved in drug metabolism, including major<br />

substrates. The circle sizes approximate relative levels <strong>of</strong> expression in human liver. The overlap <strong>of</strong><br />

the circles conveys the point that an overlap <strong>of</strong> catalytic action is <strong>of</strong>ten observed (Gibson <strong>and</strong> Skett<br />

2008, Guengerich <strong>and</strong> Montellano 2005).<br />

All three members <strong>of</strong> the CYP1 family (CYP1A2, CYP1B1, CYP1A1) are upregulated <strong>by</strong><br />

halogenated <strong>and</strong> polycyclic aromatic hydrocarbons, such as those found in cigarette smoke<br />

<strong>and</strong> charred food (e.g. benzo(a)pyrene). CYP1A2 is responsible for melatonin oxidation<br />

<strong>and</strong> uroporphyrin formation from uroporphyrinogen: it metabolizes about 24 <strong>drugs</strong> <strong>and</strong> all<br />

three members are metabolically active <strong>and</strong> detoxify many environmental carcinogens<br />

(Nelson <strong>and</strong> Nebert 2011). CYP2 is the largest P450 family in mammals with 13<br />

subfamilies <strong>and</strong> 16 genes in humans. CYP2C8, -9, -18, <strong>and</strong> -19 together metabolize more<br />

than 50 <strong>drugs</strong>, as well as arachidonic acid <strong>and</strong> steroids. CYP2D6 metabolizes more than 70<br />

<strong>drugs</strong>. CYP2A6, -A13, -B6, -E1 <strong>and</strong> -F1 are also major contributors to drug metabolism<br />

(Guengerich et al. 1998, Meyer <strong>and</strong> Zanger 1997, Nelson <strong>and</strong> Nebert 2011). P450<br />

subfamily 3A is the most abundantly expressed P450 gene in the human liver <strong>and</strong><br />

gastrointestinal tract <strong>and</strong> known to metabolize more than 120 commonly used<br />

4


INTRODUCTION<br />

pharmaceutical agents (Nelson <strong>and</strong> Nebert 2011). The human fetus is capable <strong>of</strong><br />

metabolizing a large number <strong>of</strong> compounds, <strong>and</strong> its major P450 is a member <strong>of</strong> the 3A<br />

subfamily, specifically 3A7. This single P450 comprises 30 to 50% <strong>of</strong> the total fetal P450s.<br />

However, 3A7 has not been detected in the liver <strong>of</strong> adults (Wrighton <strong>and</strong> Stevens 1992).<br />

The CYP4 family metabolizes some <strong>drugs</strong>, but is involved mainly in the metabolism <strong>of</strong><br />

fatty acids (Nelson <strong>and</strong> Nebert 2011).<br />

1.1.3 Pharmaceutical <strong>drugs</strong> development <strong>and</strong> studies<br />

In the course <strong>of</strong> modern drug development, there are four areas in which issues pertaining<br />

to drug metabolites need to be addressed, namely (i) drug metabolites as components <strong>of</strong><br />

(plasma) drug assays, (ii) drug metabolites in bioavailability studies, (iii) drug metabolites<br />

in bioequivalence studies, <strong>and</strong> (iv) drug metabolites in safety testing (Baillie et al. 2002).<br />

Pharmaceutical companies <strong>and</strong> regulatory agencies are becoming increasingly interested in<br />

the role <strong>of</strong> metabolites as potential mediators <strong>of</strong> the toxicity <strong>of</strong> new drug products (Baillie<br />

et al. 2002). In 2005, the U.S. Food <strong>and</strong> Drug Administration issued a draft guidance<br />

document "Safety testing <strong>of</strong> drug metabolites" which was followed in 2008 <strong>by</strong> the final<br />

version (Baillie 2009, Center for Drug Evaluation <strong>and</strong> Research (HFD-130) <strong>and</strong> Food <strong>and</strong><br />

Drug Administration 2008). Drugs entering the body undergo biotransformation via Phase<br />

I <strong>and</strong> Phase II metabolism. Based on the nature <strong>of</strong> the chemical reactions involved,<br />

metabolites formed from Phase I reactions can be chemically more reactive or even<br />

represent the actually bioactive form <strong>of</strong> the drug. An active metabolite may bind to the<br />

therapeutic target receptors or <strong>other</strong> receptors, interact with <strong>other</strong> targets (e.g. enzymes,<br />

proteins), <strong>and</strong> cause unintended effects (Center for Drug Evaluation <strong>and</strong> Research (HFD-<br />

130) <strong>and</strong> Food <strong>and</strong> Drug Administration 2008). It is recommended that - while there is a<br />

necessity to maintain a flexible, case-<strong>by</strong>-case evaluation <strong>of</strong> the need to perform additional<br />

safety studies on human drug metabolites - those metabolites present at disproportionately<br />

higher levels in human plasma than in any <strong>of</strong> the animal test species should be considered<br />

for safety assessment (Baillie 2009).<br />

By the early 1990s, much <strong>of</strong> the interest in P450 research had shifted to the human P450<br />

enzymes because <strong>of</strong> the availability <strong>of</strong> systems for h<strong>and</strong>ling them. In particular, studies in<br />

the areas <strong>of</strong> drug metabolism <strong>and</strong> chemical toxicology/ carcinogenesis were facilitated <strong>by</strong><br />

the knowledge that a relatively small number <strong>of</strong> the P450s account for a large fraction <strong>of</strong><br />

the metabolism <strong>of</strong> the <strong>drugs</strong>. In the pharmaceutical industry, the roles <strong>of</strong> the major hepatic<br />

P450s are extensively studied to develop predictions about bioavailability, drug-drug<br />

interactions, <strong>and</strong> toxicity (Guengerich <strong>and</strong> Montellano 2005). The metabolism <strong>of</strong> foreign<br />

5


INTRODUCTION<br />

compounds <strong>by</strong> P450s generally leads to products that are less toxic <strong>and</strong> more easily<br />

excreted. However, in some cases oxidative metabolism produces substances that are<br />

reactive toxins <strong>and</strong> mutagens (Williams et al. 2000). Toxic compounds may be detoxified<br />

following P450-catalyzed biotransformation while inert xenobiotics, including <strong>drugs</strong>, may<br />

be activated <strong>and</strong> thus become toxicants. Most xenobiotic compounds require metabolic<br />

activation <strong>by</strong> P450s to form ultimate carcinogens or toxicants. Most hepatic P450s are also<br />

present in extrahepatic tissues, <strong>and</strong> <strong>of</strong>ten at higher levels. Furthermore, some P450s are<br />

expressed preferentially in extrahepatic tissues, which may lead to unique extrahepatic<br />

metabolites <strong>and</strong> tissue-specific consequences in cellular toxicology <strong>and</strong> organ pathology.<br />

Risk assessment <strong>of</strong> potential human toxicants is currently based primarily on data obtained<br />

from animal bioassays <strong>and</strong> the knowledge <strong>of</strong> mechanism <strong>of</strong> toxicity derived from animal<br />

experiments (Ding <strong>and</strong> Kaminsky 2003). Metabolic studies using biological models are<br />

important for determining the significance <strong>of</strong> specific enzymes or pathways for toxicity<br />

(Ding <strong>and</strong> Kaminsky 2003).<br />

Pharmaceutical products (PPs) <strong>and</strong> <strong>by</strong>-products (BPs) are spread into 24 therapeutic<br />

classes, among which four are dominant in terms <strong>of</strong> references. About 40% <strong>of</strong> studies<br />

concern non-steroidal anti-inflammatory <strong>drugs</strong> (NSAIDs), the three <strong>other</strong>s being<br />

anticonvulsants, antibiotics, <strong>and</strong> lipid regulators. NSAIDs are essentially represented <strong>by</strong><br />

ibupr<strong>of</strong>en <strong>and</strong> dicl<strong>of</strong>enac. The most studied lipid regulator is gemfibrozil. Carbamazepine<br />

is the most tested anticonvulsant (Mompelat et al. 2009).<br />

1.1.4 Environmental aspects<br />

In the European Union today, about 3,000 different substances are being used in medicines<br />

such as painkillers, antibiotics, contraceptives, β-blockers, lipid regulators, tranquilizers,<br />

<strong>and</strong> impotence <strong>drugs</strong> (Ternes et al. 2004, Wiegel et al. 2004). In Germany, in the year<br />

2000, about 7,000 t <strong>of</strong> synthetic medicines for human consumption were sold. (report from<br />

the Federal Environmental Authority; Umweltbundesamt). The figure for veterinary <strong>drugs</strong><br />

was 2,320 t in 2002 (Wiegel et al. 2004). Table 1 shows some examples <strong>of</strong> pharmaceutical<br />

<strong>drugs</strong> chosen as environmentally relevant together with their annual consumption <strong>and</strong><br />

quantities remaining as unconverted substances. Human-use medicines are designed to<br />

have a biological effect <strong>and</strong> to be bioavailable. However, it is only in recent years that<br />

there has been increasing concern over the trace amounts <strong>of</strong> <strong>pharmaceuticals</strong> that are<br />

appearing in the environment <strong>and</strong> the effects they may cause (Daughton <strong>and</strong> Ternes 1999,<br />

Whitacre et al. 2010).<br />

6


INTRODUCTION<br />

Table 1 Consumption <strong>of</strong> pharmaceutical <strong>drugs</strong> in Germany in 2001 <strong>and</strong> percentages remaining<br />

unconverted (Sachverständigen Rat für Umweltfragen 2007).<br />

Pharmaceutical class/drug<br />

Consumption per year Unconverted substrate<br />

Analgesics<br />

(kg)<br />

(% <strong>of</strong> dose)<br />

Metamizole 64,400 0<br />

Phenazone 24,850 5<br />

Propyphenazone 24,180 2<br />

Codeine 9,700 70<br />

Morphine 880 10<br />

NSAIDs (see 1.1)<br />

Ibupr<strong>of</strong>en 344,880 1<br />

Dicl<strong>of</strong>enac 85,800 70<br />

Indometacin 3,700 30<br />

Ketopr<strong>of</strong>en 1,613 10<br />

Piroxicam 724 5<br />

Mecl<strong>of</strong>enamic acid not determined not determined<br />

Antitussives/cough medicines<br />

Ambroxol 14,470 6<br />

Codeine 9,700 70<br />

Dihydrocodeine 1,245 40<br />

Hydrocodone 8 40<br />

Bronchodilators/antiasthmatic<br />

<strong>drugs</strong><br />

Salbutamol 414 80<br />

Terbutaline 118 60<br />

Fenoterol 72 50<br />

Clenbuterol 1 90<br />

Antibiotics<br />

Sulfamethoxazole 53,600 33<br />

Doxycycline 24,180 2<br />

Cipr<strong>of</strong>loxacin 17,973 45<br />

Clindamycin 16,100 30<br />

Trimethoprim 11,426 80<br />

7


INTRODUCTION<br />

Roxithromycin 9,550 60<br />

Clarithromycin 7,159 35<br />

Norfloxacin 4,724 70<br />

Ofloxacin 2,280 95<br />

Oxytetracycline 2,020 not determined<br />

Tetracycline 1,530 not determined<br />

Spiramycin 300 25<br />

Chloramphenicol 202 10<br />

Chlortetracycline 99 not determined<br />

Antihypertensives/α/β-blockers<br />

Metoprolol 93,000 10<br />

Sotalol 26,600 90<br />

Atenolol 13,500 90<br />

Propranolol 3,900 5<br />

Bisoprolol 2,914 50<br />

Antiepileptics<br />

Carbamazepine 87,600 30<br />

Psychoactive <strong>drugs</strong><br />

Diazepam 1,107 30<br />

Chem<strong>other</strong>apy agents<br />

Cyclophosphamide 385 7<br />

Ifosfamide 170 50<br />

Hormones/contraceptives<br />

17α-Ethinylestradiol 50 85<br />

Data origin: SRU/Stellungnahme Nr.12-2007; Huschek <strong>and</strong> Krengel 2003<br />

The occurrence <strong>of</strong> anthropogenic organic contaminants in aquatic ecosystems has drawn<br />

the attention <strong>of</strong> both the public <strong>and</strong> scientists in recent years, a fact confirmed <strong>by</strong> the<br />

numerous studies published in this field. With respect to <strong>pharmaceuticals</strong>, many<br />

investigations have focused on high consumption substances currently being introduced<br />

into the environment (Hass et al. 2012). Accurate assessment <strong>of</strong> <strong>drugs</strong> impact on the<br />

environment is difficult, since there is a multitude <strong>of</strong> input sources in the environment with<br />

no evident quantitative data available concerning the relative distribution <strong>of</strong><br />

<strong>pharmaceuticals</strong> from all emission sources (shown in figure 3).<br />

8


INTRODUCTION<br />

Figure 3 Origin <strong>and</strong> routes <strong>of</strong> pharmaceutical products (adapted from Petrović et al. (Petrović et al.<br />

2003) <strong>and</strong> from Mompelat et al. (Mompelat et al. 2009). DWT: drinking water treatments; WWTP:<br />

waste water treatment plant.<br />

However, humans <strong>and</strong> animals account for the main contamination source <strong>of</strong> potable water<br />

resources (Mompelat et al. 2009). Once ingested <strong>and</strong> metabolized, <strong>pharmaceuticals</strong> (native<br />

<strong>and</strong>/or metabolites) are excreted via urine <strong>and</strong> feces, or unused or expired <strong>drugs</strong> are<br />

improperly disposed <strong>of</strong> in toilets, <strong>and</strong> travel in urban wastewater to the wastewater<br />

treatment plant (WWTP) or, for countryside households, they are directly released into<br />

septic tanks (Carrara et al. 2008, Mompelat et al. 2009). To a minor but relevant extent,<br />

hospital wastewater also contributes to the total loads. Pharmaceuticals can also enter<br />

agricultural l<strong>and</strong> when digested sludge is applied as fertilizer. Veterinary <strong>pharmaceuticals</strong><br />

are excreted directly into the environment, <strong>and</strong> thus contaminate soil while the resulting<br />

run<strong>of</strong>f gets into rivers. Therefore, the efficiency with which WWTPs can remove<br />

<strong>pharmaceuticals</strong> is crucial (Ternes et al. 2004).<br />

In recent decades, more than 100 different <strong>drugs</strong> have been detected in aquatic<br />

environments at concentrations from ng L -1 to µg L -1 (Daughton <strong>and</strong> Ternes 1999). Recent<br />

studies have demonstrated that, despite relatively low concentrations <strong>of</strong> <strong>pharmaceuticals</strong> in<br />

the environment (typically at sub-parts-per-billion level; < ppb), <strong>pharmaceuticals</strong> are <strong>of</strong><br />

ecological concern due to their potential long-term adverse effects on humans <strong>and</strong> wildlife<br />

(Ankley et al. 2007, Celiz et al. 2009). In most <strong>of</strong> the investigations reported to date, the<br />

efficiency <strong>of</strong> <strong>drugs</strong> removal during water treatment is determined <strong>by</strong> measuring the<br />

disappearance <strong>of</strong> the parent compound, but not the formation <strong>of</strong> <strong>by</strong>-products. Similarly,<br />

9


INTRODUCTION<br />

there is a lack <strong>of</strong> available information on the environmental fate <strong>and</strong> occurrence <strong>of</strong><br />

excreted human <strong>and</strong> animal drug metabolites. The European Medicines Agency has set<br />

guidelines for reporting concentrations <strong>of</strong> <strong>drugs</strong> <strong>and</strong> their metabolites. A systematic<br />

approach has been developed to predict the total risk associated with <strong>pharmaceuticals</strong><br />

entering the environment. The guidelines say that any metabolite formed at a concentration<br />

greater than 10 percent <strong>of</strong> the parent compound needs further investigation (Boxall et al.<br />

2004, Celiz et al. 2009).<br />

Ecotoxicity studies in the laboratory have demonstrated effects <strong>of</strong> <strong>pharmaceuticals</strong> on end<br />

points such as reproduction, growth, behavior <strong>and</strong> feeding for fish <strong>and</strong> invertebrates<br />

(Martinović et al. 2007, Quinn et al. 2008, Whitacre et al. 2010). Drugs have already been<br />

detected in fish tissues (Brooks et al. 2005). In the terrestrial environment, the catastrophic<br />

decline <strong>of</strong> Indian vulture populations was found to be the result <strong>of</strong> exposure to the human<br />

anti-inflammatory drug dicl<strong>of</strong>enac (Oaks et al. 2004, Whitacre et al. 2010). In the river<br />

Elbe, in 1998, dicl<strong>of</strong>enac, ibupr<strong>of</strong>en <strong>and</strong> carbamazepine were found as well as various<br />

antibiotics <strong>and</strong> lipid regulators in a concentration range from


INTRODUCTION<br />

terminology P450 is uncommon for enzymes because it is not based on function, but<br />

originally describes the spectral properties <strong>of</strong> this b-type heme containing red pigments,<br />

which display a typical absorption b<strong>and</strong> around 450 nm in their reduced carbon-monoxide<br />

bound form (Omura <strong>and</strong> Sato 1964). All P450s contain a cysteine-ligated heme (iron<br />

protoporphyrine IX) that can bind dioxygen (O 2 ) in the active site (Dawson <strong>and</strong> Sono<br />

1987).<br />

The <strong>of</strong>ficial nomenclature in enzymology, which has been in use since the early 1960s,<br />

follows the EC 1 classification system that is strictly based on the reaction catalyzed.<br />

Enzymes catalyzing a particular reaction get so-called EC-numbers that comprise four<br />

digits (Nomenclature Committee <strong>of</strong> the International Union <strong>of</strong> Biochemistry <strong>and</strong> Molecular<br />

Biology (NC-IUBMB) 2012).<br />

Most P450s belong to the main class <strong>of</strong> oxidoreductases (EC 1.x), more precisely to those<br />

oxidoreductases acting on paired donors, with incorporation or reduction <strong>of</strong> molecular<br />

oxygen (EC 1.14.x). There are some exceptions, where P450s are found, for example, in<br />

the main class <strong>of</strong> isomerases <strong>and</strong> lyases (intramolecular oxidoreductases, EC 5.3.x <strong>and</strong><br />

carbon-oxygen lyases, EC 4.2.x) (Degtyarenko 2004). Members <strong>of</strong> the P450 enzyme<br />

family are involved in the biotransformation <strong>of</strong> <strong>drugs</strong>, the bioconversion <strong>of</strong> xenobiotics, the<br />

metabolism <strong>of</strong> chemical carcinogens, the biosynthesis <strong>of</strong> physiologically important<br />

compounds (such as steroids, fatty acids, eicosanoids, fat-soluble vitamins, <strong>and</strong> bile acids),<br />

the conversion <strong>of</strong> alkanes, terpenes, <strong>and</strong> aromatic compounds as well as in the degradation<br />

<strong>of</strong> herbicides <strong>and</strong> insecticides (Bernhardt 2006). The reactions catalyzed can be extremely<br />

diverse <strong>and</strong> include e.g. hydroxylations, N-, O- <strong>and</strong> S-dealkylations, sulphoxidations,<br />

epoxidations, deaminations, desulphurations, dehalogenations, peroxidations, N-oxide<br />

reductions, dehydrations <strong>and</strong> isomerizations(Bernhardt 2006, Guengerich 2001, Mansuy<br />

1998). Many P450s convert a broad range <strong>of</strong> substrates <strong>and</strong> catalyze multiple reactions.<br />

For example, unspecific monooxygenases (EC 1.14.14.1) act on xenobiotics, steroids, fatty<br />

acids, vitamins <strong>and</strong> prostagl<strong>and</strong>ins <strong>and</strong> they catalyze amongst <strong>other</strong>s hydroxylations,<br />

epoxidations, N-oxidations, sulfooxidations, N-, S- <strong>and</strong> O-dealkylations, desulfations,<br />

deaminations, <strong>and</strong> the reduction <strong>of</strong> azo-, nitro- <strong>and</strong> N-oxide groups.<br />

Since many <strong>of</strong> the individual P450s catalyze multiple reactions, the usual method <strong>of</strong><br />

naming enzymes is inadequate for this group <strong>of</strong> proteins, so a systematic nomenclature has<br />

been developed on the basis <strong>of</strong> structural <strong>and</strong> sequence homology (Nelson et al., Nelson et<br />

1 EC - Enzyme Commission: Nomenclature Committee <strong>of</strong> the International Union <strong>of</strong> Biochemistry <strong>and</strong><br />

Molecular Biology (NC-IUBMB), in consultation with the IUPAC-IUBMB Joint Commission on<br />

Biochemical Nomenclature (JCBN), http://www.chem.qmul.ac.uk/iubmb/enzyme/<br />

11


INTRODUCTION<br />

al. 1996). P450 genes are identified <strong>by</strong> the abbreviation CYP followed <strong>by</strong> a number<br />

denoting the family; members <strong>of</strong> a family show more than 40% sequence identity.<br />

Following is a letter designating a subfamily whose members share more than 55%<br />

sequence identity. Finally, the second number indicates the individual gene within the<br />

subfamily (e.g. CYP2A1) (Hannemann et al. 2007). At present, there is a total <strong>of</strong> over<br />

1,000 P450 families <strong>and</strong> over 2,500 subfamilies (Nelson 2011). Altogether, more than<br />

12,000 P450 genes have been reported across all kingdoms <strong>of</strong> life. Mammals have usually<br />

between 50 <strong>and</strong> 100 encoded P450 enzymes. There are even more CYP genes found in<br />

plant genomes than in animal genomes, because plants are sessile <strong>and</strong> carry out their<br />

biochemistry de novo from carbon dioxide, ammonia, water <strong>and</strong> minerals.<br />

Table 2 Named cytochrome P450s in March 2010, adapted from Nelson et al., 2011 (Nelson 2011)<br />

Kingdom/ organisms P450 number CYP families a Sequenced<br />

genomes b<br />

Animals (Metazoa) 4088<br />

CYPs/genome<br />

Insects 2137 67 11 37-178 c<br />

Vertebrates 1461 18 8 57-102 d<br />

Invertebrates (without<br />

490 71 4 76-89 e<br />

insects)<br />

Plants (Plantae) 4267 126 14 10-332 f<br />

Fungi (Eumycota) 2784 399 53 1-153 g<br />

Protists 247 62 22 1-55 h<br />

Eubacteria 1042 333 212 1-52 i<br />

Archaea 26 13 17 1-4<br />

Viruses (Mimivirus) 2 2 1 2<br />

Total 12,456<br />

a includes all named CYP families; not<br />

just those from sequenced genomes.<br />

b<br />

only genomes included in the<br />

nomenclature in 2010<br />

c<br />

Aedes aegypti<br />

d<br />

Mouse (Mus<br />

musculus)<br />

e<br />

Tetranychus<br />

urticae<br />

f Rice (Oryza<br />

sativa)<br />

g<br />

Aspergillus<br />

oryzae<br />

h Dictyostelium<br />

discoideum<br />

i<br />

Frankia sp.<br />

Angiosperm plants may have between 250 (Arabidopsis spp.) <strong>and</strong> over 400 P450s (potato,<br />

Solanum tuberosum) that are involved in stress responses, reaction to wounding, seedling<br />

development, taste, colour <strong>and</strong> aroma <strong>of</strong> fruits <strong>and</strong> flowers, structural support (lignin),<br />

water barriers (cutins <strong>and</strong> suberins), hormone signaling <strong>and</strong> synthesis <strong>of</strong> allelochemicals<br />

12


INTRODUCTION<br />

for defence against predators (Nelson <strong>and</strong> Nebert 2011). Fungi can also possess numerous<br />

P450s; thus in the genome <strong>of</strong> the basidiomycetous white-rot fungus Phanerochaete<br />

chrysosporium, more than 150 P450 genes (most <strong>of</strong> them monooxygenases) were found<br />

(Subramanian et al. 2010), <strong>and</strong> 153 genes are present in the ascomycetous mold<br />

Aspergillus oryzae (Nelson 2011). Table 2 gives an overview <strong>of</strong> the number <strong>and</strong><br />

distribution <strong>of</strong> P450s in different groups <strong>of</strong> organisms.<br />

1.2.1 Classification <strong>of</strong> electron transport systems <strong>of</strong> P450s<br />

After the discovery <strong>of</strong> cytochrome P450s, two main classes reflecting the redox partners<br />

involved were described: the adrenal mitochondrial P450 systems obtaining electrons from<br />

NAD(P)H via adrenodoxin reductase <strong>and</strong> adrenodoxin (Hannemann et al. 2007, Omura et<br />

al. 1966), <strong>and</strong> the liver microsomal P450s obtaining electrons from NAD(P)H via a flavincontaining<br />

(FAD, FMN) P450 reductase (Hannemann et al. 2007, Lu <strong>and</strong> Coon 1968). It<br />

was not until the 1980s, when CYP102 (P450BM3) was discovered, that an example came<br />

to light showing an unusual property, the fusion <strong>of</strong> the P450 <strong>and</strong> reductase components<br />

into one polypeptide chain (Narhi <strong>and</strong> Fulco 1987). Since then, several more <strong>of</strong> such<br />

“unusual” electron transfer chains have been described for different P450s (table 3). Class<br />

II cytochrome P450s are the most common type found in all eukaryotic organisms (Črešnar<br />

<strong>and</strong> Petrič).<br />

Table 3: Classes <strong>of</strong> P450 systems grouped according to the topology <strong>of</strong> protein components<br />

involved in the electron transfer towards the actual P450 component, including schematic<br />

organization <strong>of</strong> the proteins (Hannemann et al. 2007). Protein components contain the following<br />

redox centers: Fdx (iron-sulfur-cluster); FdR, Ferredoxin reductase (FAD); CPR, cytochrome P450<br />

reductase (FAD, FMN); Fldx, Flavodoxin (FMN); OFOR, 2-oxoacid: ferredoxin oxidoreductase<br />

(thiamin pyrophosphate, [4Fe-4S] cluster); PFOR, phthalate-family oxygenase reductase (FMN,<br />

[2Fe-2S] cluster).<br />

Class/ source<br />

Class I<br />

Bacterial<br />

Mitochondrial<br />

Electron transport<br />

chain<br />

NAD(P)H►[FdR] ►[Fdx]<br />

►[P450]<br />

NADPH►[FdR] ►[Fdx]<br />

►[P450]<br />

Scheme<br />

Localization/<br />

remarks<br />

Cytosolic, soluble<br />

P450s; inner<br />

mitochondrial<br />

membrane; FdRmembrane<br />

associated; Fdxmitochondrial<br />

matrix, soluble<br />

13


INTRODUCTION<br />

Class II<br />

Bacterial,<br />

<strong>fungal</strong>,<br />

Microsomal A<br />

Microsomal B,<br />

<strong>fungal</strong><br />

Microsomal C<br />

NADPH►[CPR] ►[P450]<br />

NADPH►[CPR] ►[P450]<br />

NADPH►[CPR] ►[cytb5]<br />

► [P450]<br />

NADPH►[cytb5Red] ►<br />

[cytb5] ► [P450]<br />

Cytosolic, soluble<br />

Membrane<br />

anchored,<br />

endoplasmatic<br />

reticulum<br />

(A,B,C)<br />

Class III<br />

Bacterial<br />

NAD(P)H►[FdR] ►[Fldx]<br />

►[P450]<br />

Cytosolic,<br />

soluble,<br />

Citrobacter<br />

braakii<br />

Class IV<br />

Bacterial<br />

Pyruvate, CoA►[OFOR]<br />

►[Fdx] ►[P450]<br />

Cytosolic,<br />

soluble,<br />

Sulfolobus<br />

tokadaii<br />

Class V<br />

Bacterial<br />

NADPH►[FdR] ►[Fdx-<br />

P450]<br />

Cytosolic,<br />

soluble,<br />

Methylococcus<br />

capsulatus<br />

Class VI<br />

Bacterial<br />

NAD(P)H►[FdR] ►[Fldx-<br />

P450]<br />

Cytosolic soluble,<br />

Rhodococcus<br />

rhodochrous<br />

strain 11Y<br />

Class VII<br />

Bacterial<br />

NADPH►[PFOR-P450]<br />

Cytosolic,<br />

soluble,<br />

Rhodococcus sp<br />

strain NCIMB<br />

9784,<br />

Burkholderia sp.,<br />

Ralstonia<br />

metallidurans<br />

14


INTRODUCTION<br />

Class VIII<br />

Bacterial,<br />

<strong>fungal</strong><br />

Class IX<br />

Only NADH<br />

dependent,<br />

<strong>fungal</strong><br />

Class X<br />

Independent in<br />

plants/<br />

mammals<br />

NADPH►[CPR-P450]<br />

NADH►[P450]<br />

[P450]<br />

Cytosolic,<br />

soluble, Bacillus<br />

megaterium,<br />

Fusarium<br />

oxysporum<br />

Cytosolic,<br />

soluble, Fusarium<br />

oxysporum<br />

Membrane bound,<br />

ER, example:<br />

Thromboxane<br />

synthase<br />

1.2.2 Bacterial P450s<br />

Bacterial P450s form a large proportion <strong>of</strong> the superfamily. They are soluble enzymes<br />

whereas their eukaryotic counterparts are usually membrane-bound. This fact simplifies<br />

the overexpression <strong>and</strong> purification <strong>of</strong> bacterial enzymes (Munro <strong>and</strong> Lindsay 1996). Most<br />

eubacterial species contain at least one P450 enzyme while some archaea seemingly do not<br />

possess any P450 enzyme. It is possible that P450s may have evolved during the advent <strong>of</strong><br />

atmospheric oxygen in the atmosphere (about 1.5 to 1 billion years ago), where their<br />

original role was in the detoxification <strong>of</strong> molecular oxygen (O 2 ) in anaerobic bacteria<br />

(Lewis <strong>and</strong> Wiseman 2005, Wickramasinghe <strong>and</strong> Villee 1975).<br />

A bacterial P450 was first found in nitrogen-fixing Rhizobium bacteroides (Apple<strong>by</strong> 1967).<br />

In contrast to the microsomal <strong>and</strong> mitochondrial P450s reported before, the Rhizobium<br />

P450 was soluble (i.e. dissolved in the cytosol); however, the function was not elucidated.<br />

Then P450s in Pseudomonas putida <strong>and</strong> Coryebacterium sp. were reported (Cardini <strong>and</strong><br />

Jurtshuk 1968, Katagiri et al. 1968) . P450 cam (CYP101) from Pseudomonas was found to<br />

catalyze the NADH-dependent oxidation <strong>of</strong> camphor <strong>and</strong> required soluble NADH,<br />

putidaredoxin <strong>and</strong> NADH-putidaredoxin reductase for the reaction. Corynebacterium P450<br />

catalyzes the NADH-dependent terminal oxidation <strong>of</strong> n-octane to 1-octanol. More soluble<br />

P450s were found in various prokaryotic organisms in the following years (Omura 2005).<br />

The first self-sufficient P450-reductase fusion enzyme was found in Bacillus megaterium<br />

<strong>and</strong> named P450 BM-3 (CYP102A2). Similar P450-reductase fusion proteins were later<br />

found in <strong>other</strong> bacterial species (Narhi <strong>and</strong> Fulco 1987, Omura 2005). P450 eryF from<br />

15


INTRODUCTION<br />

Saccharopolyspora erythrea is the only bacterial P450 to which a biosynthetic function has<br />

been assigned. The bacterium produces erythromycin A, <strong>and</strong> P450 eryF catalyzes the<br />

hydroxylation <strong>of</strong> 6-deoxyerythronolide B to generate erythronolide B, which is a precursor<br />

<strong>of</strong> this macrolide antibiotic (Munro <strong>and</strong> Lindsay 1996). Among the P450s characterized<br />

from bacteria <strong>of</strong> the order Actinomycetales (Rhodococcus spp. strains) are enzymes<br />

catalyzing alkane hydroxylation, pentachlorophenol dehalogenation <strong>and</strong> alkoxyphenol O-<br />

dealkylation (Karlson et al. 1993, Munro <strong>and</strong> Lindsay 1996). The most known function <strong>of</strong><br />

bacterial P450s is the oxidation <strong>of</strong> recalcitrant organic compounds so that the hosts can<br />

utilize them as sole carbon source for growth. This is also the generally recognized<br />

difference between prokaryotic <strong>and</strong> eukaryotic P450s (involvement in carbon metabolism<br />

vs. detoxification/biosynthesis) (Fulco 1991). However, the role <strong>of</strong> bacterial P450s in the<br />

utilization <strong>of</strong> unusual organic compounds is non-essential for the viability <strong>of</strong> the<br />

organisms. The ability <strong>of</strong> bacterial P450s to metabolize compounds such as terpenes <strong>and</strong><br />

alkanes has attracted interest, not only regarding the decomposition <strong>of</strong> organic wastes <strong>and</strong><br />

pollutants, but also for biotransformations in the synthesis <strong>of</strong> <strong>pharmaceuticals</strong> <strong>and</strong> fine<br />

chemicals (Munro <strong>and</strong> Lindsay 1996). Aside from the inherent interest in the physiological<br />

roles <strong>and</strong> biotechnological potential <strong>of</strong> bacterial P450s, these enzymes have become<br />

attractive <strong>and</strong> productive targets for spectroscopic <strong>and</strong> structural analyses (Munro <strong>and</strong><br />

Lindsay 1996).<br />

1.2.3 Plant P450s<br />

Plants utilize a diverse array <strong>of</strong> P450s in their biosynthetic <strong>and</strong> detoxification pathways.<br />

Biosynthetic P450s play paramount roles in the synthesis <strong>of</strong> lignin intermediates, sterols,<br />

terpenes, flavonoids/is<strong>of</strong>lavonoids, furanocoumarins, cyanogenic glycosides, lipids, plant<br />

growth regulators (phytohormones) such as gibberellins, jasminoic acid, <strong>and</strong><br />

brassinosteroids, <strong>and</strong> a variety <strong>of</strong> <strong>other</strong> secondary plant metabolites (Chapple 1998,<br />

Nielsen et al. 2005, Schuler 1996a).<br />

Many scientific studies have shown the important role <strong>of</strong> P450s in the biosynthesis <strong>and</strong><br />

degradation <strong>of</strong> compounds important in plant insect interactions (Schuler 1996b), <strong>and</strong> the<br />

importance <strong>of</strong> plant P450s in the metabolism <strong>of</strong> herbicides (Frear 1995). Plant P450<br />

monooxygenases are membrane-bound heme proteins which consist <strong>of</strong> a protoporphyrin<br />

IX <strong>and</strong> an apoprotein that confers the substrate specificity. Plant P450s have been shown to<br />

carry out hydroxylations, epoxidations, heteroatom dealkylations <strong>and</strong> oxygenations<br />

(Bolwell et al. 1994). Cinnamic acid 4-hydroxylase from Helianthus tuberosus (Jerusalem<br />

artichoke) was the first plant P450 to be functionally characterized. This enzyme catalyzes<br />

16


INTRODUCTION<br />

an essential step in the phenylpropanoid pathway <strong>and</strong> is considered to be ubiquitous in<br />

plants (Nielsen et al. 2005). Like vertebrates <strong>and</strong> fungi, certain plants produce<br />

polyhydroxylated steroidal hormones, called brassinosteroids, to regulate <strong>and</strong> control tissue<br />

morphology. They are nonessential phytohormones with impacts on morphological<br />

characteristics, for example, leaf shape <strong>and</strong> dwarfism (Clouse <strong>and</strong> Sasse 1998, Nielsen et<br />

al. 2005).<br />

1.2.4 Fungal P450s<br />

Cytochrome P450 monooxygenases <strong>of</strong> fungi are involved in many essential cellular<br />

processes <strong>and</strong> play diverse roles. They catalyze the conversion <strong>of</strong> hydrophobic<br />

intermediates <strong>of</strong> primary <strong>and</strong> secondary metabolic pathways, detoxify natural toxins <strong>and</strong><br />

environmental pollutants <strong>and</strong> allow fungi to grow under different conditions. Fungal P450s<br />

represent three electron transport systems, Class II, Class VIII <strong>and</strong> Class IX (table 1;<br />

1.2.1). Class II P450s are involved in metabolite synthesis as well as in detoxification<br />

<strong>and</strong>/or degradation <strong>of</strong> xenobiotic compounds. Class VIII is represented <strong>by</strong> P450 foxy (from<br />

Fusarium oxysporum). This enzyme is responsible for the hydroxylation <strong>of</strong> fatty acids.<br />

P450s belonging to Class IX <strong>of</strong> electron transfer systems are involved in the process <strong>of</strong><br />

<strong>fungal</strong> denitrification (Črešnar <strong>and</strong> Petrič, Shoun et al.). The key enzyme, nitric oxide<br />

reductase (EC 1.7.1.14), is also designated as P450 nor <strong>and</strong> shows structural similarity to<br />

"classic" P450 monooxygenases. But on the <strong>other</strong> h<strong>and</strong>, P450 nor functionally differs from<br />

the rest <strong>of</strong> P450s as it catalyzes the reduction <strong>of</strong> two molecules <strong>of</strong> nitric oxide (NO) to<br />

nitrous oxide (N 2 O) where<strong>by</strong> the electrons are directly transferred from NAD(P)H to the<br />

enzyme without the need <strong>of</strong> any additional redox partner. P450 nor is the only eukaryotic<br />

P450 that is soluble, i.e. not membrane-associated (Črešnar <strong>and</strong> Petrič, Shoun et al.).<br />

An<strong>other</strong> <strong>fungal</strong> P450 family (CYP53) uses e.g. benzoic acid as substrate to produce 4-<br />

hydroxybenzoic acid or catalyzes the demethylation <strong>of</strong> 3-methoxybenzoic to 3-<br />

hydroxybenzoic acid, <strong>and</strong> has thus an essential role in the detoxification <strong>of</strong> aromatic<br />

compounds (Lah et al.). One well-studied class <strong>of</strong> bioconversions carried out <strong>by</strong> <strong>fungal</strong><br />

P450 enzymes is the (stereo)-specific hydroxylation <strong>of</strong> pharmaceutically interesting<br />

steroids, especially progesterone (Mahato <strong>and</strong> Majumdar 1993, van den Brink et al. 1998).<br />

An<strong>other</strong> P450-catalyzed process is the 14α-demethylation <strong>of</strong> lanosterol in yeasts <strong>and</strong> <strong>of</strong><br />

eburicol in filamentous fungi. Ergosterol (the product <strong>of</strong> this reaction), has a similar<br />

structure as mammalian cholesterol <strong>and</strong> performs an important role in membrane<br />

stability/fluidity <strong>and</strong> in completion <strong>of</strong> the cell cycle (van den Brink et al. 1998).<br />

17


INTRODUCTION<br />

The evolutionarily conserved (CYP51) <strong>and</strong> fungi-specific (CYP61 <strong>and</strong> CYP 56) P450s are<br />

essential for primary metabolism, enabling these microorganisms to preserve cell wall<br />

integrity <strong>and</strong> to form the spore outer wall, respectively, regardless <strong>of</strong> the fungus’<br />

morphology or ecological niche. Furthermore, several <strong>other</strong> <strong>fungal</strong> P450s with <strong>other</strong><br />

activities <strong>and</strong> functions have been characterized. They were found to catalyze e.g. the<br />

detoxification <strong>of</strong> xenobiotic compounds, the initial step in n-alkane assimilation, fatty acid<br />

hydroxylation, nitrite reduction <strong>and</strong> mycotoxin production (Črešnar <strong>and</strong> Petrič). A wellstudied<br />

P450-catalyzed reaction in fungi is the alkane assimilation/degradation <strong>by</strong> some<br />

C<strong>and</strong>ida species (Scheller et al. 1996, van den Brink et al. 1998). Table 4 below lists<br />

selected <strong>fungal</strong> P450s <strong>and</strong> their functions (those known so far).<br />

Table 4 Compilation <strong>of</strong> most known <strong>fungal</strong> P450s <strong>and</strong> their functions.<br />

Fungus<br />

(<strong>fungal</strong> group)<br />

Aspergillus<br />

parasiticus<br />

(ascomycete)<br />

C<strong>and</strong>ida albicans<br />

(ascomycetous<br />

yeast)<br />

C<strong>and</strong>ida maltosa<br />

(ascomycetous<br />

yeast)<br />

Cunninghamella<br />

bainieri<br />

(zygomycete)<br />

Curvularia lunata<br />

(ascomycete)<br />

Fusarium<br />

oxysporum<br />

(ascomycete)<br />

Fusarium<br />

sporotrichioides<br />

(ascomycete)<br />

Giberella fujikuroi<br />

(ascomycete)<br />

Nectria<br />

haematococca<br />

(ascomycete)<br />

Penicillium paxilli<br />

(ascomycete)<br />

Phanerochaete<br />

chrysosporium<br />

(basidiomycete)<br />

P450 /known<br />

information<br />

Catalyzed<br />

reaction<br />

Known<br />

substrate(s)<br />

Ref.<br />

CYP64 family oxygenation O-methylsterigmatocystin,<br />

(Yu et al. 1998)<br />

dihydrosterigmatocystin<br />

P450 14αdm (CYP51 O-demethylation steroids (Joseph-Horne<br />

family)<br />

<strong>and</strong> Hollomon<br />

CYP52 family<br />

P450<br />

oxygenation/<br />

hydroxylation<br />

hydroxylation<br />

N-demethylation<br />

n-alkane<br />

hydrocarbons<br />

codeine<br />

1997)<br />

(Ohkuma et al.<br />

1998)<br />

(Ferris et al.<br />

1976, Gibson et<br />

al. 1984)<br />

P450 lun 11β-hydroxylation steroids (Suzuki et al.<br />

1993)<br />

P450 foxy (CYP505) hydroxylation fatty acids (Nakayama et<br />

al. 1996)<br />

CYP65A1 hydroxylation mycotoxin<br />

biosynthesis<br />

CYP68 family hydroxylation kaurenoic acid<br />

(giberellin<br />

biosynthesis)<br />

(Alex<strong>and</strong>er et al.<br />

1998)<br />

(Malonek et al.<br />

2005, Tudzynski<br />

et al. 2001)<br />

CYP57 family O-demethylation pisatin (Desjardins <strong>and</strong><br />

VanEtten 1986,<br />

Maloney <strong>and</strong><br />

VanEtten 1994)<br />

P450 paxP,<br />

P450 paxQ<br />

oxygenation paspalin (paxilline<br />

biosynthesis)<br />

149 P450s estimated,<br />

12 families,<br />

23 subfamilies.<br />

Functions like plant<br />

CYP73 <strong>and</strong> CYP98<br />

Functions like bacterial<br />

CYP102 (P450 BM3 ).<br />

Seven sequences in the<br />

hydroxylation<br />

cinnamic acid<br />

(Saikia et al.<br />

2007)<br />

(Doddapaneni et<br />

al. 2005,<br />

Matsuzaki <strong>and</strong><br />

Wariishi 2004,<br />

Syed <strong>and</strong> Yadav<br />

2012)<br />

hydroxylation cumene (Matsuzaki <strong>and</strong><br />

Wariishi 2004)<br />

18


INTRODUCTION<br />

Pleurotus ostreatus<br />

(basidiomycete)<br />

Pleurotus<br />

pulmonarius<br />

(basidiomycete)<br />

Rhisopus nigricans<br />

(zygomycete)<br />

Saccharomyces<br />

cerevisiae<br />

(ascomycetous<br />

yeast). Only three<br />

P450 sequences:<br />

same phylogenic cluster<br />

as CYP102<br />

Functions like yeast<br />

CYP52, seven sequences<br />

were categorized into the<br />

CYP52 family<br />

Functions like bacterial<br />

CYP105A1 <strong>and</strong><br />

CYP105B1, mammalian<br />

CYP2B, plant CYP76B1<br />

<strong>and</strong> <strong>fungal</strong><br />

CYP510A1,but similarity<br />

above 35% only with<br />

CYP510A1<br />

Functions like<br />

mammalian CYP2A, but<br />

no sequence similarity<br />

above 35%<br />

Functions like <strong>fungal</strong><br />

CYP53, one sequence <strong>of</strong><br />

P450 from<br />

P.chrysosporium fell into<br />

the CYP53 familiy<br />

Functions like bacterial<br />

CYP101(P450 cam ), no<br />

similarity above 35%<br />

Functions like bacterial<br />

CYP176A1(P450 cin ) <strong>and</strong><br />

mammalian CYP3A, but<br />

no sequential similarity<br />

above 35%<br />

PcCYP65a2, <strong>and</strong><br />

PcCYP11a3 clones,<br />

functions like human <strong>and</strong><br />

hydroxylation 1,12-dodecanediol (Matsuzaki <strong>and</strong><br />

Wariishi 2004)<br />

deethylation<br />

Hydroxylation<br />

7-position, <strong>by</strong><br />

P.chrysosporium<br />

Four different<br />

positions<br />

hydroxylation<br />

4-position, <strong>by</strong><br />

P.chrysosporium<br />

three different<br />

positions<br />

Hydroxylation<br />

5- <strong>and</strong> 6- position,<br />

<strong>by</strong><br />

P.chrysosporium<br />

three different<br />

positions<br />

4-ethoxybenzoic<br />

acid <strong>and</strong> 7-<br />

ethoxycoumarin<br />

coumarin<br />

benzoic acid<br />

camphor<br />

(Matsuzaki <strong>and</strong><br />

Wariishi 2004)<br />

(Matsuzaki <strong>and</strong><br />

Wariishi 2004)<br />

(Matsuzaki <strong>and</strong><br />

Wariishi 2004)<br />

(Matsuzaki <strong>and</strong><br />

Wariishi 2004)<br />

hydroxylation 1,8-cineol (Matsuzaki <strong>and</strong><br />

Wariishi 2004)<br />

hydroxylation<br />

mono- <strong>and</strong><br />

dichloro-p-dioxins<br />

(Kasai et al.)<br />

rat CYP1A2<br />

CYP5136A3 hydroxylation alkylphenols, PAH (Syed et al.<br />

2011)<br />

P450 dNIR , P450 nor denitrification nitrite (Shoun et al.,<br />

Shoun <strong>and</strong><br />

Tanimoto 1991)<br />

CYP53 (cloned<br />

PcCYP1f)<br />

P450 hydroxylation pyrene,<br />

anthracene,<br />

fluorene,<br />

hydroxylation benzoate (Matsuzaki <strong>and</strong><br />

Wariishi 2005)<br />

(Bezalel et al.<br />

1996)<br />

dibenzothiophene<br />

P450 hydroxylation benzo[a]pyrene (Masaphy et al.<br />

1995, Maspahy<br />

et al. 1999)<br />

P450 11α hydroxylation steroids (Bossche <strong>and</strong><br />

Koymans 1998,<br />

Breskvar et al.<br />

1995)<br />

CYP61 desaturation steroids (Kelly et al.<br />

2005, Nelson<br />

2009)<br />

19


INTRODUCTION<br />

CYP51F1,CYP56A1<br />

<strong>and</strong> CYP61A1<br />

Trametes (Coriolus)<br />

versicolor<br />

(basidiomycete)<br />

Trichosporon<br />

moniliiforme<br />

<strong>and</strong> T.cutaneum<br />

(basidiomycetous<br />

yeast)<br />

Yarrowia lipolytica<br />

(ascomycetous<br />

yeast)<br />

Cloned gene, similarity<br />

with CYP512A1, but less<br />

then 40% identity with<br />

any P450<br />

sulfur-oxidizing<br />

reactions<br />

DBT derivatives<br />

(Ichinose et al.<br />

2002)<br />

P450 decarboxylation salicylate, phenol (Stundl et al.<br />

2000)<br />

CYP52 family<br />

oxygenation/<br />

hydroxylation<br />

n-alkane (Iida et al. 2000)<br />

1.2.5 Animal P450s<br />

1.2.5.1 Insect P450s<br />

In insects, the number <strong>of</strong> cytochrome P450s in sequenced genomes ranges from 37 in the<br />

body louse, Pediculus humanus (Lee et al. 2010, Sztal et al. 2012), up to 160 (178,<br />

corrected in 2011 (Nelson 2011)) in the dengue mosquito, Aedes aegypti (Strode et al.<br />

2008, Sztal et al. 2012). There are four large clades <strong>of</strong> insect P450 genes that existed<br />

before the divergence <strong>of</strong> the class <strong>of</strong> Insecta <strong>and</strong> that are also present in vertebrates: the<br />

CYP2 clade, the CYP3 clade, the CYP4 clade <strong>and</strong> the mitochondrial P450 clade.<br />

Mitochondrial P450s are not found in plants or fungi, but seem to be restricted to animals<br />

(Feyereisen 2006). P450s from each <strong>of</strong> these clades have been linked to insecticide<br />

resistance or to the metabolism <strong>of</strong> natural products <strong>and</strong> xenobiotics. In particular, insects<br />

appear to maintain a repertoire <strong>of</strong> mitochondrial P450 paralogues devoted to the response<br />

to environmental stress (Feyereisen 2006). The metabolism <strong>of</strong> xenobiotics, insecticides in<br />

particular, to less toxic metabolites is probably the best-known feature <strong>of</strong> P450s<br />

(Feyereisen 1999). Cytochrome P450 monooxygenases are involved in many cases <strong>of</strong><br />

resistance <strong>of</strong> insects to insecticides. Resistance has been proposed to be associated with<br />

increase in monooxygenase activities <strong>and</strong> the cytochrome P450 content (Bergé et al. 1998).<br />

Insect P450s can metabolize a wide range <strong>of</strong> plant allelochemicals, including<br />

furanocoumarins, terpenoids, indoles, glucosinolates, flavonoids, cardenolides,<br />

phenylpropenes, ketohydrocarbons, alkaloids, lignans, pyrethrins, <strong>and</strong> the is<strong>of</strong>lavonoid<br />

rotenone (Li et al. 2007). However, there is an increasing number <strong>of</strong> studies indicating that<br />

P450s have a much broader role in insects than just detoxification. For example,<br />

biosynthesis <strong>of</strong> juvenile <strong>and</strong> molting hormones are processes in which the role <strong>of</strong> insect<br />

P450s has been well documented (Scott 2008).<br />

20


INTRODUCTION<br />

1.2.5.2 Mammalian/ Human P450s<br />

There are eight P450 families in mammalians, which are involved in the metabolism <strong>of</strong><br />

<strong>pharmaceuticals</strong> <strong>and</strong> <strong>drugs</strong>, products <strong>of</strong> incomplete combustion (e.g. polycyclic aromatic<br />

hydrocarbons, PAHs) as well as <strong>of</strong> plant ingredients (e.g. phenolics, phytoalexins). It has<br />

been proposed that four <strong>of</strong> these P450 gene families (I, II, III, <strong>and</strong> IV) evolved <strong>and</strong><br />

diverged in animals due to their exposure to plant ingredients <strong>and</strong> decayed plant products<br />

during the last one billion years. The overlapping substrate specificities <strong>of</strong> these P450<br />

enzymes are remarkable (Nebert <strong>and</strong> Gonzalez 1987). There are at least 43 gene families<br />

recognized in animals. In humans, 57 P450s were identified <strong>and</strong> this number seems to be<br />

complete because <strong>of</strong> our excellent knowledge <strong>of</strong> the human genome (Guengerich <strong>and</strong><br />

Montellano 2005, Nelson 2009).<br />

Most eukaryotic P450s are intrinsic membrane proteins that are expressed in the<br />

endoplasmic reticula <strong>of</strong> plant, <strong>fungal</strong>, <strong>and</strong> animal cells. Animal cells also express P450s<br />

that are located in the inner membrane <strong>of</strong> mitochondria. The human enzymes contribute to<br />

the synthesis <strong>of</strong> steroid hormones, prostacyclin, thromboxane, cholesterol, <strong>and</strong> bile acids,<br />

as well as to the degradation <strong>of</strong> endogenous compounds, such as fatty acids, retinoic acid,<br />

<strong>and</strong> steroids, <strong>and</strong> exogenous compounds that include <strong>drugs</strong> <strong>and</strong> carcinogens (Williams et<br />

al. 2000). In humans, <strong>of</strong> the P450s with known catalytic activities, 14 are involved in<br />

steroidogenesis, 4 are involved in certains aspects <strong>of</strong> vitamin metabolism, 5 are involved in<br />

the metabolism <strong>of</strong> eicosanoids, 4 have fatty acids as substrates, <strong>and</strong> 15 catalyze the<br />

transformation <strong>of</strong> xenobiotics. Many <strong>of</strong> the xenobiotic-converting P450s can also oxidize<br />

steroids <strong>and</strong> fatty acids. Most <strong>of</strong> the steroid-oxidizing enzymes are constitutive <strong>and</strong> their<br />

levels are relatively invariable, in contrast to the xenobiotic-metabolizing P450s, the levels<br />

<strong>of</strong> which can vary considerably (Guengerich <strong>and</strong> Montellano 2005).<br />

The organ that expresses the highest levels <strong>of</strong> P450s is the liver, which plays the dominant<br />

role in the first-pass clearance <strong>of</strong> ingested xenobiotic compounds <strong>and</strong> controls the systemic<br />

level <strong>of</strong> <strong>drugs</strong> <strong>and</strong> <strong>other</strong> chemicals. Extrahepatic tissues, especially those that are the<br />

portals <strong>of</strong> entry for xenobiotic compounds, such as the respiratory <strong>and</strong> the gastrointestinal<br />

tracts, also express P450s. In these tissues, 450 enzymes not only contribute to the firstpass<br />

clearance, but may also influence the tissue burden <strong>of</strong> xenobiotic compounds or<br />

bioavailability <strong>of</strong> therapeutic agents (Ding <strong>and</strong> Kaminsky 2003).<br />

1.2.6 Engineered P450s<br />

The catalytic versatility, substrate diversity, <strong>and</strong> sheer number <strong>of</strong> P450s have led to<br />

considerable interest over the last 25 years in the engineering <strong>of</strong> P450 enzymes for a<br />

21


INTRODUCTION<br />

variety <strong>of</strong> purposes (Gillam 2008). Mammalian P450 enzymes, as membrane-bound twoenzyme<br />

systems, are notoriously difficult to express <strong>and</strong> hard to maintain in an active form.<br />

Parallel heterologous expression in addition must provide exact ratios <strong>of</strong> components to<br />

ensure catalytic activity (Vail et al. 2005). P450s can be engineered using rational <strong>and</strong><br />

evolutionary approaches. Rational approaches are characterized <strong>by</strong> the deliberate mutation<br />

<strong>of</strong> one or more amino acids based on mechanistic or structural information <strong>and</strong> are<br />

appropriate for altering P450 selectivities (Jung et al., Urlacher <strong>and</strong> Girhard). By directed<br />

evolution, mutations are introduced in a r<strong>and</strong>om or semi-r<strong>and</strong>om manner, for example <strong>by</strong><br />

site-saturation mutagenesis at residues thought to be important for the desired property,<br />

<strong>and</strong> the resulting mutant P450s are screened for enhancement <strong>of</strong> that property or set <strong>of</strong><br />

properties. A recent area <strong>of</strong> interest is the design <strong>and</strong> development <strong>of</strong> novel P450s as<br />

biocatalysts <strong>of</strong> reactions <strong>of</strong> commercial interest, such as in pharmaceutical or fine chemical<br />

synthesis (Gillam 2008).<br />

For example, over the past decade, P450 BM3 was engineered for the oxidation <strong>of</strong> alkanes,<br />

terpenes, heteroaromatics, alkaloids, steroids <strong>and</strong> <strong>other</strong> chemical substances, which are<br />

either hydroxylated, epoxidized or demethylated (Jung et al., Urlacher <strong>and</strong> Girhard). P450<br />

BM3 from Bacillus megaterium (discovered in the 1980s (Warman et al. 2005)) possesses<br />

favorable properties that makes it an attractive target for engieneering (Munro et al. 2002).<br />

It is a soluble bacterial, highly active enzyme, which is naturally fused to its reductase <strong>and</strong><br />

routinely expressed in E. coli. The hydroxylation <strong>of</strong> arachidonic acid catalyzed <strong>by</strong> P450<br />

BM3 is the most rapid P450-catalyzed hydroxylation known so far (k cat = 17,000 min -1 )<br />

(Jung et al., Warman et al. 2005).<br />

In the 1990s the technique <strong>of</strong> DNA shuffling (Crameri et al. 1998, Stemmer 1994),<br />

involving homology-dependent recombination <strong>of</strong> fragments <strong>of</strong> orthologous P450 genes in a<br />

primerless PCR reassembly, was applied for P450-engineering with great success (Gillam<br />

2008). A novel approach for DNA recombination (the so-called SCHEMA algorithm) was<br />

used to obtain chimeras based on P450 BM3 <strong>and</strong> its homologs CYP102A2 <strong>and</strong> CYP102A3<br />

from Bacillus subtilis. A survey <strong>of</strong> the activities <strong>of</strong> new chimeras demonstrated that this<br />

approach created completely new functions absent in the wild-type enzymes, including the<br />

ability to oxidize <strong>drugs</strong> (Urlacher <strong>and</strong> Girhard).<br />

1.2.7 Catalyzed reactions<br />

P450 enzymes play two main roles in living organisms. Some cytochrome P450s catalyze<br />

oxidation steps involved in the biosynthesis or biodegradation <strong>of</strong> endogenous compounds<br />

like steroids, fatty acids or prostagl<strong>and</strong>ins. The second group is responsible for oxidative<br />

22


INTRODUCTION<br />

biotransformation <strong>of</strong> exogenous substances from the environment, facilitating their<br />

elimination from living organisms (Mansuy 1998). Cytochrome P450s catalyze more than<br />

20 different reaction types (Sono et al. 1996). Some <strong>of</strong> the "atypical" reactions are<br />

oxidations involving C-C or C=N bond cleavage, reductions, dehydrations,<br />

dehydrogenations or isomerizations. The classical reactions performed <strong>by</strong> P450s involve<br />

the transfer <strong>of</strong> one oxygen atom from dioxygen (O 2 ) to the substrate.<br />

R<br />

+<br />

+<br />

− H + O2 + NAD( P)<br />

H + H → R − OH + H<br />

2O<br />

+ NAD(<br />

P)<br />

1.2.8 The most important P450 reactions are summarized in Table 5.<br />

Reaction mechanism <strong>of</strong> P450s<br />

The catalytic cycle <strong>of</strong> P450 has been studied over decades. The active center for catalysis<br />

is the iron(III)-protoporphyrin IX (heme) with the thiolate <strong>of</strong> the conserved cysteine<br />

residue as fifth lig<strong>and</strong> (Werck-Reichhart <strong>and</strong> Feyereisen 2000) (Figure 4). Four pyrrole<br />

rings are bonded <strong>by</strong> methine bridges (-CH 2 =), giving a planar <strong>and</strong> highly conjugated<br />

porphyrin system (Nagababu <strong>and</strong> Rifkind 2004). Six<br />

heteroatoms coordinate the heme iron octahedrally. The<br />

four porphyrin nitrogens are the equatorial lig<strong>and</strong>s. The<br />

remaining two axial lig<strong>and</strong>s are located below (fifth<br />

coordination site, proximal) <strong>and</strong> above (sixth coordination<br />

site, distal) the plane <strong>of</strong> the heme (Cirino <strong>and</strong> Arnold<br />

2003).<br />

Figure 4 Iron (III) protoporphyrin-IX linked with a proximal cysteine lig<strong>and</strong>. Modified according to<br />

Meunier et al. (Meunier et al. 2004).<br />

The P450 reaction cycle has been proposed <strong>by</strong> several authors(Guengerich 2001, Munro et<br />

al. 2007, Ogliaro et al. 2002) <strong>and</strong> can be summarized as follows (Figure 5). It is initiated<br />

<strong>by</strong> the binding <strong>of</strong> a substrate, usually with concomitant displacement <strong>of</strong> the distal water<br />

lig<strong>and</strong> [(a), H 2 O…heme(Fe III )] (Ortiz de Montellano et al. 2005, Sligar <strong>and</strong> Gunsalus<br />

1976). The substrate binds to the species (a) near the distal region <strong>of</strong> the heme, which<br />

causes a dehydration <strong>of</strong> the active site resulting in a five-coordinated heme iron within the<br />

enzyme-substrate complex [(b), R-H…heme(Fe III )].<br />

23


MATERIALS AND METHODS<br />

24<br />

Table 5 Oxidative <strong>and</strong> non-oxidative reactions catalyzed <strong>by</strong> P450 enzymes(Guengerich 2001, Isin <strong>and</strong> Guengerich 2007, Mansuy 1998, Sono et al. 1996)<br />

INTRODUCTION


INTRODUCTION<br />

25


INTRODUCTION<br />

MATERIALS AND METHODS<br />

26


INTRODUCTION<br />

The latter accepts a first electron that is derived from NAD(P)H <strong>and</strong> supplied via<br />

flavin/ferredoxin or diflavin reductases to form the ferrous state <strong>of</strong> the enzyme [(c), (R-<br />

H)…heme(Fe II )] (Groves 2006, Ullrich <strong>and</strong> H<strong>of</strong>richter 2007). The ferrous complex is<br />

reactive enough to add dioxygen (O 2 ) resulting in the formation <strong>of</strong> a ferrous dioxycomplex<br />

[(d), R-H…heme(Fe II )-O 2 ], which then accepts the second electron (Groves 2006,<br />

Ullrich <strong>and</strong> H<strong>of</strong>richter 2007). This results in formation <strong>of</strong> a ferric peroxy ion [(e), (R-<br />

H)…heme(Fe III -O - 2 )], which is protonated to form the hydroperoxy complex (also referred<br />

as Compound 0, [(f), (R-H)…heme(Fe III -O-OH)]. Compound 0 undergoes heterolytic<br />

cleavage between the oxygen atoms giving rise to the putative oxo-ferryl state [(g), (R-<br />

H)…heme(Fe IV =O) •+ ], referred to as Compound I (Rittle <strong>and</strong> Green 2010).<br />

Figure 5 The catalytic cycle <strong>of</strong> cytochrome P450s monooxygenases, including the peroxide “shunt”<br />

pathway. RH is the substrate <strong>and</strong> ROH the product <strong>of</strong> the reaction. The rhombus symbolizes the<br />

heme. Modified according to Ullrich <strong>and</strong> H<strong>of</strong>richter (Ullrich <strong>and</strong> H<strong>of</strong>richter 2007) <strong>and</strong> Werck-<br />

Reichhart <strong>and</strong> Feyereisen (Werck-Reichhart <strong>and</strong> Feyereisen 2000). (a) native (hydro)ferric enzyme<br />

(resting state), (b) ferric heme-substrate complex, (c) ferrous heme-substrate complex, (d) ferrousdioxygen<br />

complex, (e) ferric peroxy anion complex, (f) ferric hydroperoxy complex (Compound<br />

0), (g) putative oxy-ferryl radical complex (Compound I), (h) product-ferric enzyme complex.<br />

27


MATERIALS AND METHODS<br />

This electron-deficient complex may abstract a hydrogen atom or one electron from the<br />

substrate to form a sigma complex (g) with the substrate. A subsequent collapse <strong>of</strong> the<br />

intermediate or an intermediate pair in step (g) generates the product. (In the case <strong>of</strong> a<br />

hydrogen abstraction mechanism, step (g) is referred to as an “oxygen rebound.”) In step<br />

(h), the product dissociates from the enzyme, <strong>and</strong> the cycle can restart (Guengerich 2001,<br />

Isin <strong>and</strong> Guengerich 2007). The so called “shunt” pathway is a remarkable side reaction <strong>of</strong><br />

many P450s, in the course <strong>of</strong> which substrate is directly oxidized <strong>by</strong> hydrogen peroxide (or<br />

an organic peroxide) to the hydroperoxo-ferryl state (f) (Ullrich <strong>and</strong> H<strong>of</strong>richter 2007).<br />

1.3 Peroxygenases<br />

Since February 2011, a second sub-subclass (EC 1.11.2.x) <strong>of</strong> peroxidases (EC 1.11.x) has<br />

been listed in the Enzyme Nomenclature system. The enzymes <strong>of</strong> this sub-subclass are<br />

named peroxygenases <strong>and</strong> catalyze reactions with a peroxide as acceptor, one oxygen atom<br />

<strong>of</strong> which is incorporated into the product<br />

(http://www.chem.qmul.ac.uk/iubmb/enzyme/EC1/11/2/). This means peroxygenases<br />

catalyze the direct transfer <strong>of</strong> one oxygen atom from a peroxide to the substrate (→ monoperoxygenation).<br />

The peroxide (preferably hydrogen peroxide, H 2 O 2 ) is reduced <strong>and</strong> the<br />

substrate is oxidized (Hanano et al. 2006):<br />

RH + H<br />

2<br />

O2<br />

→ ROH + H<br />

2O<br />

On a functional level, peroxygenases can be regarded as hybrids <strong>of</strong> cytochrome P450<br />

monooxygenases <strong>and</strong> "classic" heme peroxidases.<br />

The sub-subclass <strong>of</strong> peroxygenases contains four entries: EC 1.11.2.1 (unspecific<br />

peroxygenase, the key enzyme <strong>of</strong> this group), EC 1.11.2.2 (myeloperoxidase), EC 1.11.2.3<br />

(plant seed peroxygenase) <strong>and</strong> EC 1.11.2.4 (fatty-acid peroxygenase). However, it has to<br />

be mentioned that in the <strong>other</strong> sub-subclass (EC 1.11.1: peroxidases) an enzyme is listed -<br />

chloroperoxidase (CfuCPO, EC 1.11.1.10) - that catalyzes, in addition to halogenation<br />

reactions, also a number <strong>of</strong> peroxygenase-type reactions (Allain et al. 1993, Omura 2005).<br />

CfuCPO, which was already discovered 50 years ago [(Hager et al. 1966), belongs as do<br />

the P450s to the heme-thiolate proteins <strong>and</strong> is apparently produced only <strong>by</strong> the<br />

ascomycetous sooty mold Caldariomyces fumago (Omura 2005).<br />

28


INTRODUCTION<br />

All unspecific peroxygenases 2 characterized so far are also heme-thiolate proteins<br />

secreted <strong>by</strong> agaric basidiomycetes (colloquially called mushrooms). They are very stable,<br />

highly glycosylated proteins that have been shown to [per]oxygenate <strong>and</strong> oxidize various<br />

organic substances (for details, see 1.3.1 below). Myeloperoxidase (MPO; until 2011,<br />

listed under EC 1.11.1.7) contains calcium <strong>and</strong> a covalently bound heme with histidine as<br />

proximal lig<strong>and</strong> (Booth et al. 1989, Davey <strong>and</strong> Fenna 1996). It is present in<br />

mammalian/human phagosomes <strong>of</strong> neutrophils <strong>and</strong> monocytes (Kleban<strong>of</strong>f 1999). MPO<br />

catalyzes the hydrogen peroxide mediated peroxygenation <strong>of</strong> halide ions (chloride,<br />

bromide, iodide) <strong>and</strong> <strong>of</strong> the pseudohalide thiocyanate, according to the following reaction<br />

(Harrison <strong>and</strong> Schultz 1976):<br />

H<br />

2<br />

O<br />

2<br />

X = Cl<br />

+ X<br />

−<br />

−<br />

, Br<br />

+ H O<br />

−<br />

, I<br />

−<br />

3<br />

+<br />

, SCN<br />

→ HOX<br />

−<br />

+ 2H<br />

2<br />

O<br />

Products <strong>of</strong> this reaction (hypohalous acids) <strong>and</strong> their following products are strong<br />

oxidants <strong>and</strong> responsible for killing phagocytized bacteria <strong>and</strong> viruses. MPO differs from<br />

CPO (EC 1.11.1.10) in its preference for formation <strong>of</strong> free hypochlorite over the<br />

chlorination <strong>of</strong> organic substrates under physiological conditions (pH 5-8). Hypochlorite in<br />

turn forms a number <strong>of</strong> antimicrobial products (Cl 2 , chloramines, hydroxyl radical, <strong>and</strong><br />

singlet oxygen). In the absence <strong>of</strong> halides, MPO can oxidize phenols <strong>and</strong> exhibits a<br />

moderate peroxygenase activity towards styrene (Fiedler et al. 2000, Kleban<strong>of</strong>f 2005,<br />

Tuynman et al. 2000).<br />

Plant seed peroxygenase is a heme protein with a calcium binding motif <strong>of</strong> the caleosintype.<br />

Enzymes <strong>of</strong> this type include membrane-bound proteins found in seeds <strong>of</strong> different<br />

plants. It catalyzes the direct transfer <strong>of</strong> one oxygen atom from an organic hydroperoxide<br />

(R-OOH), which is reduced to the corresponding alcohol, to a substrate that will be<br />

oxidized. Reactions catalyzed include hydroxylation, epoxidation <strong>and</strong> sulfoxidation.<br />

Preferred substrate <strong>and</strong> co-substrate are unsaturated fatty acids <strong>and</strong> fatty acid<br />

hydroperoxides, respectively. Plant seed peroxygenase is thought to be involved in the<br />

synthesis <strong>of</strong> cutin (Hanano et al. 2006, Lequeu et al. 2003).<br />

Fatty acid peroxygenase is a bacterial, cytosolic heme-thiolate protein with sequence<br />

homology to P450 monooxygenases. Unlike the latter, it needs neither NAD(P)H,<br />

2 In earlier publications, unspecific peroxygenase was also referred to as haloperoxidase, haloperoxidaseperoxygenase,<br />

mushroom/<strong>fungal</strong> peroxygenase, AaP (Agrocybe aegerita peroxidase/peroxygenase), CrP<br />

(Coprinellus radians peroxygenase) or APO (aromatic peroxygenase). In February 2011, these enzymes were<br />

classified as unspecific peroxygenase in the Enzyme Nomenclature system (EC 1.11.2.1; systematic name:<br />

substrate:hydrogen peroxide oxidoreductase, RH-hydroxylating or -epoxidizing; for more details see<br />

following web page: http://www.chem.qmul.ac.uk/iubmb/enzyme/EC1/11/2/1.html).<br />

29


MATERIALS AND METHODS<br />

dioxygen nor specific reductases for function. Enzymes <strong>of</strong> this type are produced <strong>by</strong><br />

eubacteria (e.g. Sphingomonas paucimobilis, P450 SPα <strong>and</strong> Bacillus subtilis, P450 BSβ ) <strong>and</strong><br />

catalyze the hydroxylation <strong>of</strong> long chain fatty acids (e.g. myristic acid) using hydrogen<br />

peroxide as oxidant/electron acceptor to produce hydroxylated fatty acids (Imai et al. 2000,<br />

Lee et al. 2003, Shoji et al. 2010).<br />

Caldariomyces fumago chloroperoxidase (CfuCPO) catalyzes the unspecific chlorination,<br />

bromination, <strong>and</strong> iodation <strong>of</strong> a variety <strong>of</strong> electrophilic organic substances via hypohalous<br />

acids as actual halogenating agent. In the absence <strong>of</strong> halide, CfuCPO mimics<br />

peroxygenases <strong>and</strong> P450s <strong>by</strong> catalyzing allylic <strong>and</strong> benzylic hydroxylations, alcohol<br />

oxidations, heteroatom dealkylations, <strong>and</strong> oxo transfer to alkenes, alkynes, alkyl- <strong>and</strong><br />

arylamines, <strong>and</strong> sulfides; however, CfuCPO does not peroxygenate aromatic rings or<br />

alkanes (H<strong>of</strong>richter <strong>and</strong> Ullrich 2006, H<strong>of</strong>richter et al. 2010, Hrycay <strong>and</strong> B<strong>and</strong>iera 2012).<br />

1.3.1 Fungal unspecific peroxygenases<br />

Like CfuCPO, <strong>fungal</strong> peroxygenases are functional hybrids <strong>of</strong> P450 monooxygenases <strong>and</strong><br />

heme-peroxidases, but with a more pronounced oxygen transfer activity. Agrocybe<br />

aegerita aromatic peroxygenase (AaeAPO) was the first enzyme <strong>of</strong> this type discovered<br />

in 2004 <strong>and</strong> is, for the time being, the best-characterized aromatic/unspecific<br />

peroxygenase. At first, this enzyme was reported to oxidize aryl alcohols to the<br />

corresponding aldehydes <strong>and</strong> hence to the corresponding benzoic acids. AaeAPO was<br />

found to catalyze also the conversion <strong>of</strong> typical peroxidase substrates, such as ABTS (2,2’-<br />

azinobis-(3-ethylbenzothiazoline-6-sulfonate) or DMP (2,6-dimethoxyphenol);<br />

furthermore, it brominates MCD (monochlordimedon) <strong>and</strong> phenol (Ullrich et al. 2004).<br />

Later it turned out that the enzyme epoxidizes/hydroxylates toluene <strong>and</strong> naphthalene over a<br />

wide pH range (3-10) with an optimum around pH 7 (Kluge et al. 2007, Ullrich <strong>and</strong><br />

H<strong>of</strong>richter 2005). Further studies showed that AaeAPO has a very broad reaction spectrum<br />

that includes, for example, sulfoxidations (e.g. <strong>of</strong> dibenzothiophene, thioanisole), N-<br />

oxidations (e.g. <strong>of</strong> pyridine), the O-dealkylation <strong>of</strong> diverse ethers (e.g. <strong>of</strong> tetrahydr<strong>of</strong>uran,<br />

diethyl ether, alkyl aryl ethers), the O-demethylenation <strong>of</strong> benzodioxoles, N-dealkylations<br />

(e.g. <strong>of</strong> secondary <strong>and</strong> tertiary amines), the epoxidation/hydroxylation <strong>of</strong> polycyclic<br />

aromatic hydrocarbons (PAHs), aliphatic hydroxylations (e.g. <strong>of</strong> n-alkanes) <strong>and</strong> the<br />

epoxidation <strong>of</strong> alkenes (Ar<strong>and</strong>a et al. 2009a, Ar<strong>and</strong>a et al. 2009b, Kinne et al. 2009a,<br />

Kinne et al. 2009b, Kinne et al. 2008, Kinne et al. 2010, Kluge et al. 2007, Kluge et al.<br />

2012, Peter et al. 2011, Poraj-Kobielska et al. 2011b, Ullrich 2008, Ullrich et al. 2008,<br />

Ullrich <strong>and</strong> H<strong>of</strong>richter 2007). In the course <strong>of</strong> several studies, numerous basidiomycetes<br />

have been screened for peroxygenase acitivity. As an outcome, the second true<br />

30


INTRODUCTION<br />

peroxygenase was discovered in 2006 in the inky caps Coprinopsis verticillata (CveAPO)<br />

<strong>and</strong> Coprinellus radians (CraAPO) (Anh 2008). The latter enzyme was characterized in<br />

2007 <strong>and</strong> shown to oxidize veratryl <strong>and</strong> benzyl alcohol as well as ABTS, DMP <strong>and</strong><br />

guaiacol into the corresponding aldehydes/benzoic acids <strong>and</strong> quinones, respectively.<br />

CraAPO is able to brominate phenol <strong>and</strong> to regioselectively epoxidize/hydroxylate<br />

naphthalene to form preferably 1-naphthol (Anh et al. 2007). Further studies demonstrated<br />

that CraAPO catalyzes the sulfoxidation <strong>and</strong> hydroxylation <strong>of</strong> dibenzothiophene, the<br />

[per]oxygenation <strong>of</strong> various PAHs, as well as the O- <strong>and</strong> N-dealkylation <strong>of</strong> many<br />

<strong>pharmaceuticals</strong> (Ar<strong>and</strong>a et al. 2009a, Ar<strong>and</strong>a et al. 2009b, Poraj-Kobielska et al. 2011b).<br />

An additional interesting peroxygenase was found in the pinwheel mushroom, Marasmius<br />

rotula. This enzyme (MroAPO), which is smaller than the <strong>other</strong> peroxygenases (32 kDa<br />

instead <strong>of</strong> 40-45 kDa), can oxidize all typical peroxidase <strong>and</strong> peroxygenase substrates such<br />

as ABTS, DMP, veratryl <strong>and</strong> benzyl alcohols, naphthalene, <strong>and</strong> pyridine. Furthermore, it<br />

hydroxylates <strong>and</strong> O-/N- dealkylates many <strong>pharmaceuticals</strong> including steroids that are not<br />

converted <strong>by</strong> the <strong>other</strong> peroxygenases (Gröbe et al. 2012, Yarman et al. 2012) (Poraj-<br />

Kobielska, 2012 unpublished results).<br />

The full-length sequences <strong>of</strong> the peroxygenases <strong>of</strong> A. aegerita, C. radians <strong>and</strong> M. rotula<br />

have recently been obtained. The sequences <strong>of</strong> AaeAPO, CraAPO <strong>and</strong> MroAPO do not<br />

show any homology to classic heme peroxidases (such as lignin, manganese or horseradish<br />

peroxidase) or to cytochrome P450 monooxygenases, <strong>and</strong> display less than 30% sequence<br />

homology to CfuCPO (Pecyna et al. 2009). The CraAPO sequence shows 59% identity <strong>and</strong><br />

69% similarity to AaeAPO, <strong>and</strong> 27% identity <strong>and</strong> 42% similarity to CfuCPO. The sequence<br />

<strong>of</strong> MroAPO differs considerably from those <strong>of</strong> the <strong>other</strong> known <strong>fungal</strong> peroxygenases,<br />

showing only 33% identity <strong>and</strong> 49% similarity to AaeAPO, 29% identity <strong>and</strong> 43%<br />

similarity to CfuCPO, <strong>and</strong> 31% identity <strong>and</strong> 45% similarity to CraAPO (Pecyna 2012).<br />

Last but not least it should be mentioned that hundreds <strong>of</strong> peroxygenase-like sequences<br />

from all classes <strong>of</strong> true fungi (Eumycota: Basidio-, Asco-, Glomero-, Zygo-,<br />

Chytridiomycota) as well as from Oomycota (e.g. Phytophtora spp.) can be found in<br />

genetic data banks (H<strong>of</strong>richter et al. 2010, Pecyna et al. 2009). The significance <strong>of</strong> these<br />

putative peroxygenases remains to be investigated.<br />

1.4 Aims <strong>and</strong> objectives<br />

In recent years, increasing scientific attention has been given to <strong>pharmaceuticals</strong> <strong>and</strong> their<br />

metabolites. Both their effects on the human body <strong>and</strong> the environment are the subject <strong>of</strong><br />

physiological, toxicological <strong>and</strong> ecotoxicological research approaches, dealing in particular<br />

with the fate <strong>of</strong> <strong>drugs</strong>. To bring more light into this field, it is important to know the<br />

31


MATERIALS AND METHODS<br />

properties <strong>of</strong> drug metabolites <strong>and</strong> their effects on living organisms. On the <strong>other</strong> h<strong>and</strong>,<br />

there is a need to remove such substances from the environment in order to reduce their<br />

risk potential. In many organisms, cytochrome P450 enzymes play a key role in the<br />

conversion <strong>and</strong> detoxification <strong>of</strong> bioactive compounds including <strong>pharmaceuticals</strong>.<br />

However, P450s are intracellular biocatalysts <strong>and</strong> consequently relatively unstable. They<br />

need expensive c<strong>of</strong>actors <strong>and</strong> are difficult to purify, which hampers their broad use in an<br />

isolated form. For a few years it has been known that <strong>fungal</strong> peroxygenases can catalyze<br />

similar reactions to those catalyzed <strong>by</strong> P450s, <strong>and</strong> several studies have demonstrated their<br />

great biotechnological potential for oxyfunctionalizations. The peroxygenases are<br />

extracellular - i.e. they are actively secreted <strong>by</strong> the fungus - <strong>and</strong> highly stable, <strong>and</strong> they do<br />

not need c<strong>of</strong>actors <strong>other</strong> than hydrogen peroxide. Peroxygenases catalyze a wide variety <strong>of</strong><br />

reactions, including hydroxylations, epoxidations, <strong>and</strong> dealkylations. These catalytic<br />

properties suggest that they may also be an interesting alternative for the enzymatic<br />

conversion <strong>of</strong> <strong>pharmaceuticals</strong> both in terms <strong>of</strong> metabolite preparation <strong>and</strong> drug removal.<br />

Accordingly, my PhD work has addressed the following questions:<br />

1. Are <strong>fungal</strong> peroxygenases (APOs) capable <strong>of</strong> oxygenating complex<br />

<strong>pharmaceuticals</strong>? If so, what are the limitations in terms <strong>of</strong> substrate structure <strong>and</strong><br />

size?<br />

2. What are the products <strong>of</strong> these conversions?<br />

3. Are these products similar to human drug metabolites formed <strong>by</strong> P450s?<br />

4. Are illicit <strong>drugs</strong> <strong>and</strong> narcotics subject to peroxygenase attack?<br />

5. Do peroxygenases <strong>of</strong> different <strong>fungal</strong> origin produce different patterns <strong>of</strong><br />

metabolites?<br />

6. Are peroxygenases able to oxidize <strong>pharmaceuticals</strong> at low, environmentally<br />

relevant concentrations?<br />

7. Can peroxygenases be used in long term experiments over several days/weeks? If<br />

so, what is their performance under such conditions?<br />

Finally, on the basis <strong>of</strong> the answers to the questions posed above, this study has aimed to<br />

develop perspectives for the application <strong>of</strong> <strong>fungal</strong> peroxygenases in the pharmaceutical<br />

industry <strong>and</strong> the water treatment sector.<br />

32


MATERIALS AND METHODS<br />

2 Materials <strong>and</strong> Methods<br />

2.1 Chemicals <strong>and</strong> reactants<br />

2.1.1 Pharmaceuticals <strong>and</strong> their metabolites<br />

Metoprolol, oseltamivir phosphate, 4-hydroxytolbutamide, 4'-hydroxydicl<strong>of</strong>enac, 5-<br />

hydroxydicl<strong>of</strong>enac, 3-hydroxyacetaminophen, omeprazole, <strong>and</strong> sildenafil were obtained<br />

from Chemos GmbH (Regenstauf, Germany). 3-Hydroxycarbamazepine, 1-<br />

hydroxyibupr<strong>of</strong>en, 2-hydroxyibupr<strong>of</strong>en, 1-oxoibupr<strong>of</strong>en <strong>and</strong> N-desmethylsildenafil were<br />

purchased from Toronto Research Chemicals Inc. (Toronto, Canada), <strong>and</strong> α-<br />

hydroxymetoprolol, glycinexylidide <strong>and</strong> monoethylglycinexylidide from TLC<br />

PharmaChem. Inc. (Vaughan, Canada). 17α-Ethinylestradiol, (R)-naproxen, 4-<br />

hydroxypropranolol, 5-hydroxypropranolol, (S)-N-desisopropylpropranolol <strong>and</strong> O-<br />

desmethylmetoprolol were obtained from Fluka (St. Gallen, Switzerl<strong>and</strong>), from Shanghai<br />

FWD Chemicals Ltd. (Shanghai, China), Biomol GmbH (Hamburg, Germany), SPIbio,<br />

Bertin Group (Montigny Le Bretonneux, France), ABX Advanced Biochemical<br />

Compounds (Radeberg, Germany) <strong>and</strong> S<strong>and</strong>oo Pharmaceuticals & Chemicals Co., Ltd.<br />

(Zhejiang, China), (S)-mephenytoin, S-desmethyl mephenytoin <strong>and</strong> 4’-hydroxy<br />

mephenytoin were purchased from Santa Cruz Biotechnology Inc. (Heidelberg, Germany)<br />

respectively. H 18 2 O 2 (90 atom %, 2% wt/vol) was a product <strong>of</strong> Icon Isotopes (New York,<br />

USA). All <strong>other</strong> chemicals were purchased from Sigma-Aldrich (Schnelldorf, Germany).<br />

Phenacetin-d 1 (N-(4-[1- 2 H]ethoxyphenyl)acetamide) was prepared from acetaminophen<br />

(N-(4-hydroxyphenyl)acetamide) <strong>and</strong> ethyl iodide-d 1 as described previously for<br />

phenacetin-d 3 (Garl<strong>and</strong> et al. 1977). The reaction product <strong>of</strong> the synthesis (phenacetin-d 1 )<br />

was identified <strong>by</strong> comparison <strong>of</strong> retention time, UV absorption spectra, <strong>and</strong> mass spectra<br />

relative to authentic phenacetin (Nottebaum <strong>and</strong> McClure 2010); mass spectrum (m/z, %):<br />

180 (100, - H), 138 (62, - C 2 H 2 O), 110 (12, - C 4 DH 6 O), 109 (96, - C 3 DH 2 O 2 ), 108 (100, -<br />

C 3 DH 6 ON), 80 (16), 43 (18).<br />

2.1.2 Drugs <strong>and</strong> their metabolites<br />

Codeine hydrochloride, dextroamphetamine hydrochloride, (R,S)-Methamphetamine, 3,4-<br />

methylenedioxyamphetamine (tenamphetamine), norcodeine, benzoylecgonine, lysergic<br />

acid diethylamne, flurazepam, midazolam <strong>and</strong> Δ9-tetrahydrocannabinol were purchased<br />

from Lipomed GmbH (Weil am Rhein, Germany). All <strong>other</strong> <strong>drugs</strong> <strong>and</strong> metabolites were<br />

obtained from Sigma-Aldrich (Schnelldorf, Germany).<br />

33


MATERIALS AND METHODS<br />

2.1.3 Steroids <strong>and</strong> anabolic agents<br />

Cortisone, pregnenolone <strong>and</strong> cholesterol were purchased from TCI Deutschl<strong>and</strong> GmbH<br />

(Eschborn, Germany). Testosteron was obtained from Fisher Scientific GmbH (Schwerte,<br />

Germany). All <strong>other</strong> steroids, anabolic steroids <strong>and</strong> metabolites were obtained from Sigma-<br />

Aldrich (Schnelldorf, Germany).<br />

2.1.4 Benzodioxole derivatives<br />

Safrole, sesamol, myristicin, 3,4-methylenedioxymethamphetamine (MDMA, ecstasy) <strong>and</strong><br />

5-chloro-1,3-benzodioxole were purchased from Sigma-Aldrich (Schnelldorf, Germany).<br />

5-Nitro-1,3-benzodioxole (3,4-methylenedioxynitrobenzene, MDONB, NBD), 4-Bromo-<br />

1,2-methylenedioxybenzene (3,4-methylenedioxybromobenzene) were obtained from TCI<br />

Deutschl<strong>and</strong> GmbH (Eschborn, Germany). Pisatin was obtained from Apin Chemicals Ltd.<br />

(Abington, United Kingdom).<br />

2.2 Fungal cultivation <strong>and</strong> purification <strong>of</strong> peroxygenases<br />

2.2.1 Agrocybe aegerita <strong>and</strong> its peroxygenase<br />

Agrocybe aegerita (Black poplar mushroom) is a typical agaric fungus, whose fruiting<br />

bodies consist <strong>of</strong> a stalk <strong>and</strong> a cap with lamellae (gills) on the underside (Fig. 6; Foto <strong>by</strong><br />

René Ullrich). The fruiting bodies are edible <strong>and</strong><br />

have been cultivated in Mediterranean countries<br />

for centuries. A. aegerita preferably colonizes<br />

hardwood (Populus spp., Acer spp.) where it<br />

causes an atypical white-rot, but also mulch <strong>and</strong><br />

straw-like materials (Stamets 1993).<br />

The extracellular peroxygenase <strong>of</strong> A. aegerita<br />

(AaeAPO; is<strong>of</strong>orm II, 44 kDa) was produced <strong>and</strong><br />

purified as described previously (Ullrich et al.<br />

2004). Agrocybe aegerita (<strong>fungal</strong> strain: A1,<br />

isolated from a straw-pile in Jena/Wöllnitz in<br />

1970) was cultivated on agar plates <strong>and</strong> in liquid<br />

cultures, the culture liquid <strong>of</strong> which was<br />

ultrafiltered <strong>and</strong> concentrated.<br />

Figure 6 Agrocybe aegerita (Photo <strong>by</strong> René Ullrich)<br />

34


MATERIALS AND METHODS<br />

These crude enzyme preparations were further purified <strong>by</strong> three steps <strong>of</strong> ion exchange<br />

chromatography using the strong cation exchangers SP Sepharose FF <strong>and</strong> Mono S,<br />

<strong>and</strong> the strong anion exchanger Mono Q. All chromatographic steps were performed<br />

with an ÄKTA FPLC system (Fast Protein Liquid Chromatography, GE Healthcare<br />

Europe GmbH, Freiburg, Germany). After every exchange step, enzyme-containing<br />

fractions were collected, ultrafiltered <strong>and</strong> desalted <strong>by</strong> a 150-mL or 20-mL stirred cell<br />

system (10 kDa cut-<strong>of</strong>f modified polyethersulphone membrane, Pall Life Sciences,<br />

Dreieich, Germany). After the purification, the molecular mass <strong>of</strong> enzyme is<strong>of</strong>orms was<br />

determined <strong>by</strong> SDS-PAGE (Novex XCell; Invitrogen GmbH, Karlsruhe, Germany) using a<br />

12% NuPage gel. Protein concentration was determined with a low-molecular-mass<br />

calibration kit served as a protein st<strong>and</strong>ard (MBI Fermentas®, St. Leon Roth, Germany).<br />

The same electrophoresis system was used for analytical isoelectric focusing (IEF).<br />

Electrophoretically separated protein b<strong>and</strong>s were visualized with the Colloidal Blue<br />

Staining Kit (Invitrogen). Final enzyme preparations were sterile-filtered, portionized <strong>and</strong><br />

frozen at -80°C.<br />

2.2.1.1 Fungal cultivation<br />

Fungal stock cultures (A. aegerita strain A1; DSM 22459) were maintained on 2% maltextract<br />

agar (MEA) plates <strong>and</strong> kept in the dark at 4°C (Pecyna et al. 2009). For subsequent<br />

studies, <strong>fungal</strong> precultures were prepared <strong>by</strong> excising one agar plug (∅ 1 cm) from the<br />

stock culture, transferring it onto fresh MEA plates <strong>and</strong> keeping the cultures at 24°C for 2-<br />

3 weeks. Afterwards, 27 500-mL flasks containing 6 g <strong>of</strong> soybean flour (Hensel, Magstadt,<br />

Germany) in 200 mL <strong>of</strong> destilled water, were sterilized <strong>and</strong> inoculated with the contents <strong>of</strong><br />

one <strong>fungal</strong> preculture homogenized in 80 mL sterile water (0.05 % v/v). Fungal cultures<br />

were agitated on a rotary shaker (100 rpm) at 24°C for 2-3 weeks.<br />

2.2.1.2 Ultrafiltration<br />

The <strong>fungal</strong> mycelium was removed <strong>by</strong> filtration (filter GF6; Schleicher & Schuell, Dassel,<br />

Germany) <strong>and</strong> the enzyme-containing liquid was concentrated 60-fold <strong>by</strong> two steps <strong>of</strong><br />

ultrafiltration using a tangential-flow cassette system (Ultrasette, cut-<strong>of</strong>f 10 kDa, Pall<br />

GmbH, Distributor VWR GmbH, Darmstadt, Germany).<br />

2.2.1.3 SP Sepharose FF separation<br />

The pH <strong>of</strong> the crude enzyme preparation was adjusted to 4.75 <strong>and</strong> the proteins were<br />

separated on the strong cation exchanger SP Sepharose Fast Flow (column XK 26/20, GE<br />

35


MATERIALS AND METHODS<br />

HealthCare). Sodium acetate buffer (10 mM, pH 4.75) was used as a mobile phase <strong>and</strong> 10<br />

mM sodium acetate buffer containing 2 M sodium chloride (pH 4.75) was used as the<br />

eluent. The column was operated at a flow rate <strong>of</strong> 13 mL min -1 <strong>and</strong> fractions (10 mL) were<br />

collected. The elution <strong>of</strong> protein was performed with a linear gradient from 0% to 50%<br />

eluent in 8 column volumes (modified according to (Ullrich et al. 2004)).<br />

2.2.1.4 Mono Q separation<br />

The pH <strong>of</strong> the enzyme preparation was adjusted to 6.0, <strong>and</strong> the mixture separated <strong>by</strong> a<br />

Mono Q 10/100 GL column. Separations were carried out with sodium acetate buffer (10<br />

mM, pH 6.0) as mobile phase <strong>and</strong> sodium acetate buffer (10 mM, pH 6.0) containing<br />

sodium chloride (2 M) as eluent. Fractions (2 mL) were collected at a flow rate <strong>of</strong> 4 mL<br />

min -1 . The elution <strong>of</strong> protein was achieved with a linear gradient from 0% to 30% eluent in<br />

7 column volumes.<br />

2.2.1.5 Mono S separation<br />

The pH <strong>of</strong> the enzyme preparation was adjusted to 4.75, <strong>and</strong> the mixture separated <strong>by</strong> a<br />

Mono S 10/100 GL column. Separations were carried out with sodium acetate buffer (10<br />

mM, pH 4.75) as solvent, <strong>and</strong> sodium acetate buffer (10 mM, pH 4.75) containing sodium<br />

chloride (1 M) as eluent. Fractions (2 mL) were collected at a flow rate <strong>of</strong> 6 mL min -1 . The<br />

elution <strong>of</strong> protein was performed with a linear gradient from 0% to 50% eluent in 10<br />

column volumes.<br />

Figure 7 Elution pr<strong>of</strong>ile recorded after ion exchange on the strong cation exchanger Mono S.<br />

Separation <strong>of</strong> different AaeAPO is<strong>of</strong>orms (I, II, III, IV). For all further studies with AaeAPO, the<br />

is<strong>of</strong>orm II was used.<br />

36


MATERIALS AND METHODS<br />

2.2.1.6 Sodium dodecyl sulfate polyacrylamide gel<br />

electrophoresis<br />

(SDS-PAGE)<br />

Molecular weights <strong>of</strong> proteins were determined <strong>by</strong> sodium dodecyl sulfate (SDS)-<br />

polyacrylamide gel electrophoresis (PAGE) in a Mini-Cell system (XCell SureLock TM ,<br />

Invitrogen GmbH, Karlsruhe, Germany) with Novex® Pre-Cast gels. The method is<br />

described in the manufacturer’s instructions. After electrophoresis, the gels were stained<br />

<strong>and</strong> protein b<strong>and</strong>s were visualized with Colloidal Blue® (Invitrogen). A low molecular<br />

weight protein calibration kit (MBI Fermentas, St. Leon Roth, Germany) was used as the<br />

st<strong>and</strong>ard (Ullrich et al. 2004). Figure 8 shows exemplary electrophoretic separations: an<br />

SDS-PAGE gel <strong>and</strong> an IEF gel <strong>of</strong> purified AaeAPO.<br />

Figure 8 Separated AaeAPO is<strong>of</strong>orm II visualized with the Colloidal Blue® Staining Kit in an<br />

SDS-PAGE gel (left) <strong>and</strong> an IEF gel (right). M: marker, M W : molecular weight, pI: isoelectric<br />

point.<br />

2.2.1.7 Isoelectric focusing (IEF)<br />

IEF separation was done in a Mini-Cell system (XCell SureLock TM , Invitrogen GmbH,<br />

Karlsruhe, Germany) applying pre-cast IEF gels (pH 3-7) with IEF markers pH 3–10 <strong>and</strong><br />

37


MATERIALS AND METHODS<br />

relevant buffers (Invitrogen GmbH, Karlsruhe, Germany), according to the manufacturer’s<br />

instructions, <strong>and</strong> previous studies (Ullrich et al. 2009, Ullrich et al. 2004).<br />

2.2.1.8 Summarized properties <strong>of</strong> the AaeAPO preparation<br />

The enzyme preparation was homogeneous <strong>by</strong> SDS-PAGE, had a molecular mass <strong>of</strong> 46<br />

kDa, a pI <strong>of</strong> 5.1 <strong>and</strong> exhibited an A 418 /A 280 ratio <strong>of</strong> 1.75 (1.86 3 , 1.68 4 ). The specific<br />

activity <strong>of</strong> the AaeAPO preparation was 117 U mg -1 (65 U mg -1 3 , 85 U mg -1 4 ) where 1 U<br />

represents the oxidation <strong>of</strong> 1 µmol <strong>of</strong> 3,4-dimethoxybenzyl alcohol (veratryl alcohol) to<br />

3,4-dimethoxybenzaldehyde (veratraldehyde) in 1 min at 23 °C (Ullrich et al. 2004).<br />

2.2.2 Coprinellus radians <strong>and</strong> its peroxygenase<br />

This fungus prefers nitrogen-rich habitats <strong>and</strong> grows both on dung, straw-like materials<br />

<strong>and</strong> wet wood. It is found not only outdoors, but also indoors on wet wood in basements<br />

<strong>and</strong> bathrooms (Uljé 2001).<br />

The strain used here, DSM 888 (ATCC 20014) isolated from horse dung, was purchased<br />

from the DSMZ, Braunschweig,<br />

Germany (Deutsche Sammlung von<br />

Mikroorganismen und Zellkulturen<br />

GmbH 2012). The unspecific<br />

peroxygenase <strong>of</strong> Coprinellus radians<br />

(CraAPO) was produced <strong>and</strong> purified as<br />

described previously (Anh et al. 2007).<br />

The fungus was cultivated on agar plates<br />

<strong>and</strong> transferred to liquid cultures, <strong>and</strong><br />

the culture liquid thus obtained was<br />

ultrafiltered <strong>and</strong> concentrated.<br />

Figure 9 Fruiting bodies <strong>of</strong> Coprinellus radians (Photo <strong>by</strong> Hung Anh)<br />

Crude enzyme preparations were further purified <strong>by</strong> two steps <strong>of</strong> anion exchange<br />

chromatography (AEX) using Q Sepharose FF <strong>and</strong> Mono Q as separation media<br />

followed <strong>by</strong> size exclusion chromatography (SEC, Superdex 75). All chromatographic<br />

steps were performed with an ÄKTA FPLC system (Fast Protein Liquid<br />

Chromatography, GE Healthcare Europe GmbH, Freiburg, Germany). After every<br />

3 second purification<br />

4 Is<strong>of</strong>orm III that was used in long-time experiments (see 2.8.3)<br />

38


MATERIALS AND METHODS<br />

exchange step, enzyme fractions were collected, ultrafiltered <strong>and</strong> desalted with 150-mL or<br />

20-mL stirred cell systems (10 kDa cut-<strong>of</strong>f modified polyethersulphone membrane, Pall<br />

Life Sciences, Dreieich, Germany). After purification, the enzyme was characterized. The<br />

molecular mass <strong>of</strong> enzyme is<strong>of</strong>orms was determined <strong>by</strong> SDS-PAGE (Novex XCell;<br />

Invitrogen GmbH, Karlsruhe, Germany) using a 12% NuPage gel. Protein concentration<br />

was determined with a low-molecular-mass calibration kit (MBI Fermentas ® , St. Leon<br />

Roth, Germany). The same electrophoresis system was used for analytical isoelectric<br />

focusing (IEF). Electrophoretically separated protein b<strong>and</strong>s were visualized with the<br />

Colloidal Blue Staining Kit (Invitrogen GmbH, Karlsruhe, Germany). The final enzyme<br />

preparations were sterile-filtered, portionized <strong>and</strong> frozen at -80°C.<br />

2.2.2.1 Fungal cultivation<br />

Fungal stock cultures <strong>of</strong> C. radians (DSM 888) were cultivated on MEA plates <strong>and</strong> kept in<br />

the dark at 4°C. For subsequent studies, <strong>fungal</strong> precultures were prepared <strong>by</strong> excising one<br />

agar plug (∅ 1 cm) from the stock culture, transferring it onto fresh MEA plates <strong>and</strong><br />

keeping the cultures at 24°C for 2-3 weeks. Afterwards, 27 500-mL flasks containing 6 g<br />

soybean flour <strong>and</strong> 8 g glucose in 200 mL destilled water, were sterilized <strong>and</strong> inoculated<br />

with the contents <strong>of</strong> a <strong>fungal</strong> preculture homogenized in 80 mL sterile water (0.05% v/v).<br />

Fungal cultures were agitated on a rotary shaker (100 rpm) at 24°C for 2-3 weeks.<br />

2.2.2.2 Ultrafiltration<br />

The <strong>fungal</strong> mycelium was removed <strong>by</strong> filtration (filter GF6; Schleicher & Schuell, Dassel,<br />

Germany) <strong>and</strong> the enzyme-containing liquid was concentrated 60-fold <strong>by</strong> two steps <strong>of</strong><br />

ultrafiltration using a tangential-flow cassette system (Ultrasette, cut-<strong>of</strong>f 10 kDa, Pall<br />

GmbH, Distributor VWR GmbH, Darmstadt, Germany).<br />

2.2.2.3 Q Sepharose FF separation<br />

The pH value <strong>of</strong> crude enzyme was adjusted to 6.0, <strong>and</strong> the proteins were separated on the<br />

strong anion exchanger Q Sepharose Fast Flow (column XK 26/20, GE HealthCare) <strong>and</strong><br />

fractionated. As the mobile phase, 10 mM sodium acetate buffer (pH 6.0) was used <strong>and</strong> 10<br />

mM sodium acetate buffer with 2 M sodium chloride (pH 6.0) served as the eluent. The<br />

fraction size was 8 mL at a flow rate <strong>of</strong> 13 mL min -1 . The elution <strong>of</strong> protein was achieved<br />

<strong>by</strong> a linear gradient from 0% to 50% eluent in 10 column volumes.<br />

39


MATERIALS AND METHODS<br />

2.2.2.4 Mono Q separation<br />

The pH value <strong>of</strong> the enzyme preparation was adjusted to 5.5 <strong>and</strong> the mixture separated on a<br />

Mono Q 10/100 GL column <strong>and</strong> fractionated. Separations were carried out with sodium<br />

acetate buffer (10 mM, pH 5.5) as solvent, <strong>and</strong> sodium acetate buffer (10 mM, pH 5.5)<br />

with sodium chloride (2 M) as eluent. The fraction size was 2 mL at a flow rate <strong>of</strong> 4 mL<br />

min -1 . The elution <strong>of</strong> protein was performed with a linear gradient from 0% to 30% eluent<br />

in 7 column volumes.<br />

2.2.2.5 Size exclusion chromatography (SEC) using<br />

Superdex 75<br />

The pH value <strong>of</strong> the enzyme preparation was adjusted to 6.5, <strong>and</strong> the mixture separated <strong>by</strong><br />

a HiLoad 26/60 Superdex 75 preparative grade column <strong>and</strong> fractionated. As the mobile<br />

phase <strong>and</strong> eluent, 50 mM sodium acetate buffer (pH 6.5) with 0.1 M sodium chloride (pH<br />

6.5) was used. The fraction size was 4 mL at a flow rate <strong>of</strong> 2.5 mL min -1 . The elution <strong>of</strong><br />

proteins occurred under isocratic conditions.<br />

Figure 10 Elution pr<strong>of</strong>ile <strong>of</strong> CraAPO recorded<br />

after size exclusion chromatography on a<br />

Superdex TM 75 column.<br />

2.2.2.6 Sodium dodecyl sulfate polyacrylamide gel<br />

electrophoresis<br />

(SDS-PAGE) <strong>and</strong> isoelectric focusing (IEF)<br />

The methods used were the same as described for AaeAPO (2.2.1.6 <strong>and</strong> 2.2.1.7). Figure 11<br />

shows as an example the SDS-PAGE gel <strong>of</strong> the purified Coprinellus radians peroxygenase<br />

(CraAPO).<br />

40


MATERIALS AND METHODS<br />

Figure 11 Purified CraAPO visualized in an SDS-PAGE gel<br />

with the Colloidal Blue staining kit. M: Marker, M W :<br />

Molecular weight.<br />

2.2.2.7 Summarized properties <strong>of</strong> the CraAPO preparation<br />

The enzyme preparation was homogeneous <strong>by</strong> SDS-PAGE, had a molecular mass <strong>of</strong> 44<br />

kDa, a pI <strong>of</strong> 4.2, <strong>and</strong> exhibited an A 419 /A 280 ratio <strong>of</strong> 1.04. The specific activity was 25.8 U<br />

mg -1 , where 1 U means the oxidation <strong>of</strong> 1 µmol <strong>of</strong> 3,4-dimethoxybenzyl alcohol (veratryl<br />

alcohol) to 3,4-dimethoxybenzaldehyde (veratraldehyde) in 1 min at 23 °C (Anh et al.<br />

2007).<br />

2.2.3 Marasmius rotula <strong>and</strong> its peroxygenase<br />

Marasmius rotula (Pinwheel<br />

mushroom) grows in deciduous<br />

forests <strong>and</strong> fruits in groups or<br />

clusters on dead wood (especially<br />

beech), woody debris such as<br />

twigs or sticks, <strong>and</strong> occasionally<br />

on rotting leaves. M. rotula<br />

peroxygenase was a gift from K.<br />

Scheibner <strong>and</strong> G. Gröbe (Lausitz<br />

University <strong>of</strong> Applied Sciences,<br />

Senftenberg, Germany).<br />

Figure 12 Marasmius rotula (Foto <strong>by</strong> René Ullrich)<br />

41


MATERIALS AND METHODS<br />

Cultivation <strong>of</strong> the fungus as well as enzyme purification <strong>and</strong> characterization were<br />

described in (Gröbe et al. 2012). The enzyme had a molecular mass <strong>of</strong> 32 kDa, a pI<br />

between 4.97 <strong>and</strong> 5.27, <strong>and</strong> a specific activity <strong>of</strong> 12.8 U mg -1 .<br />

2.3 UV-Vis spectroscopy<br />

2.3.1 Veratryl alcohol assay for peroxygenases<br />

Figure 13 Veratryl<br />

alcohol assay<br />

The assay was routinely performed in quartz cuvettes with a light path <strong>of</strong> 10 mm. The<br />

reaction mixture contained in a total <strong>of</strong> 1 mL the ingredients given in Table 5. The reaction<br />

was started <strong>by</strong> addition <strong>of</strong> H 2 O 2 <strong>and</strong> followed at 310 nm <strong>and</strong> room temperature for at least<br />

20 seconds. Enzyme activity was calculated using the initial linear increase in absorbance<br />

<strong>and</strong> the extinction coefficient for veratryl aldehyde at 310 nm (ε 310 = 9,300 M -1 cm -1 ) (Tien<br />

<strong>and</strong> Kirk 1984, Ullrich et al. 2004). Note that the same assay is used to measure the activity<br />

<strong>of</strong> lignin peroxidase (LiP, EC 1.11.1.14), except that acidic tartrate buffer (pH 3.0) is used<br />

for Lignin peroxidase instead <strong>of</strong> neutral phosphate buffer (H<strong>of</strong>richter et al. 2010).<br />

Table 5 Composition <strong>of</strong> veratryl alcohol assay.<br />

Compound<br />

Solution<br />

concentration (mM)<br />

Amount (µL)<br />

End<br />

concentration (mM)<br />

Potassium phosphate 100 500 50<br />

buffer pH 7<br />

Veratryl alcohol 50 100 5<br />

H 2 O 290-389<br />

Sample/ Enzyme<br />

1-100<br />

solution<br />

H 2 O 2 100 10 1<br />

42


MATERIALS AND METHODS<br />

2.3.2 ABTS assay for laccase<br />

Figure 14 ABTS assay for laccases<br />

The assay was routinely performed in quartz cuvettes with a light path <strong>of</strong> 10 mm. The<br />

reaction mixture contained in a total <strong>of</strong> 1 mL the ingredients listed in Table 6. The reaction<br />

was started <strong>by</strong> the addition <strong>of</strong> the enzyme-containing sample to the mixture <strong>and</strong> was<br />

followed at 420 nm <strong>and</strong> room temperature for over 60 seconds. Enzyme activity was<br />

calculated using the initial linear increase in absorbance <strong>and</strong> the extinction coefficient for<br />

the ABTS cation radical at 420 nm (ε 420 = 36,000 M -1 cm -1 ) (Wolfenden <strong>and</strong> Willson<br />

1982).<br />

Table 6 Composition <strong>of</strong> the ABTS [2,2'-azino-bis(3-ethylbenzothiazoline-6-sulphonic acid)] assay.<br />

Compound<br />

Solution<br />

concentration (mM)<br />

Amount (µL)<br />

End<br />

concentration (mM)<br />

Sodium citrate buffer 100 500 50<br />

pH 4.5<br />

ABTS 6 50 0.3<br />

H 2 O 350-445<br />

Sample/Enzyme<br />

solution<br />

5-100<br />

2.3.3 Nitrobenzodioxole assay for peroxygenases<br />

2.3.3.1 Spectrophotometric activity measurement at neutral pH<br />

Figure 15 Nitrobenzodioxole assay for peroxygenases<br />

This is a new assay developed in the course <strong>of</strong> this work (see 3.10) <strong>and</strong> it has been<br />

published recently (Poraj-Kobielska et al. 2011a). Peroxygenases <strong>of</strong> Agrocybe aegerita,<br />

43


MATERIALS AND METHODS<br />

Coprinellus radians, Coprinopsis verticillata 5 , Marasmius rotula as well as recombinant<br />

peroxygenase <strong>of</strong> Coprinopsis cinerea 6 (0.05-2 Units mL -1 ; ~0.02-10 µM) were individually<br />

dissolved together with 5-nitro-1,3-benzodioxole [NBD; 1,2-(methylenedioxy)-4-<br />

nitrobenzene (0.125-0.5 mM) 7 ] in potassium phosphate buffer (50 mM, pH 7.0) containing<br />

acetonitrile (10% vol/vol) <strong>and</strong>, if the reaction products were to be quantified <strong>by</strong> HPLC,<br />

ascorbic acid 8 (4.0 mM). After adding the co-substrate H 2 O 2 (1 mM), the reaction mixture<br />

immediately changed its color to bright yellow (at pH 7 <strong>and</strong> 23 °C).<br />

The assay was routinely performed in quartz cuvettes with a light path <strong>of</strong> 10 mm. The<br />

reaction mixture contained in a total <strong>of</strong> 1 mL the ingredients given in Table 7. The reaction<br />

was started <strong>by</strong> the addition <strong>of</strong> H 2 O 2 <strong>and</strong> followed at 425 nm <strong>and</strong> room temperature for 15-<br />

30 seconds. Peroxygenase activity was calculated using the initial linear increase in<br />

absorbance <strong>and</strong> the extinction coefficient for 4-nitrocatechol at 425 nm (ε 425 = 9,700 M -1<br />

cm -1 ).<br />

Table 7 Composition <strong>of</strong> the nitrobenzodioxole (NBD) assay.<br />

Compound<br />

Potassium phosphate<br />

buffer pH 7<br />

Solution Amount (µL)<br />

End<br />

concentration (mM)<br />

concentration (mM)<br />

100 500 50<br />

NBD 50 100 5<br />

H 2 O 290-389<br />

Sample 1-100<br />

H 2 O 2 100 10 1<br />

2.3.3.2 Colorimetric detection under alkaline conditions<br />

The colorimetric detection <strong>of</strong> 4-nitrocatechol can be performed as an end-point<br />

determination <strong>and</strong> requires the adjustment <strong>of</strong> the assay solution mentioned above, after a<br />

certain reaction time (5-10 min), to a pH >12, e.g. <strong>by</strong> addition <strong>of</strong> 100 µL sodium lye (10<br />

M). In the presence <strong>of</strong> 4-nitrocatechol, the assay mixture will immediately change its color<br />

to red <strong>and</strong> peroxygenase activity can be calculated <strong>by</strong> detecting the absorbance at 514 nm<br />

(ε 514 = 11,400 M −1 cm −1 ).<br />

5 gift <strong>of</strong> E. Ar<strong>and</strong>a <strong>and</strong> R. Ullrich (International Graduate School <strong>of</strong> <strong>Zittau</strong>)<br />

6 gift <strong>of</strong> H. Lund (Novozymes A/S, Copenhagen, Denmark)<br />

7 NBD is hardly soluble in water <strong>and</strong> aqueous solvent mixtures at concentrations higher than 0.5 mM. Stock<br />

solutions can be prepared <strong>by</strong> using pure acetonitrile as solvent.<br />

8 Ascorbic acid is added to prevent the further oxidation <strong>of</strong> the product 4-nitrocatechol, which proceeds via<br />

the peroxidative activity <strong>of</strong> A. aegerita peroxygenase (Kinne et al. 2009a, Kinne et al. 2008). The addition <strong>of</strong><br />

the radical scavenger is not necessary for the qualitative colorimetric detection <strong>of</strong> peroxygenases.<br />

44


MATERIALS AND METHODS<br />

2.3.3.3 Tests with environmental samples<br />

Fresh samples <strong>of</strong> leaf-litter from a riparian Alnus forest near <strong>Zittau</strong> (Germany) as well as <strong>of</strong><br />

chopped wheat straw <strong>and</strong> beech wood were homogenized <strong>and</strong> 2 g <strong>of</strong> them extracted in<br />

potassium phosphate buffer (50 mM, pH 7.0) for 24 h. The crude extracts were centrifuged<br />

<strong>and</strong> the supernatants used for the experiment. The peroxygenase activity messurements in<br />

the extracts showed no positive results. The test was performed analogue to the<br />

nitrobenzodioxole assay (2.3.3.1) in quartz cuvettes with a light path <strong>of</strong> 10 mm. The<br />

reaction mixture contained in a total <strong>of</strong> 1 mL the ingredients given in Table 8. The reaction<br />

was started <strong>by</strong> the addition <strong>of</strong> H 2 O 2 <strong>and</strong> followed at 425 nm <strong>and</strong> room temperature for 15-<br />

30 seconds. Peroxygenase activity was calculated using the initial linear increase in<br />

absorbance <strong>and</strong> the extinction coefficient for 4-nitrocatechol at 425 nm (ε 425 = 9,700 M -1<br />

cm -1 ).<br />

Table 8 Composition <strong>of</strong> the nitrobenzodioxole (NBD) assay with environmental samples.<br />

Compound<br />

Solution<br />

concentration (mM)<br />

Amount (µL)<br />

End<br />

concentration (mM)<br />

Potassium phosphate 100 500 50<br />

buffer pH 7<br />

NBD 50 100 5<br />

extract 300<br />

H 2 O 70<br />

AaeAPO 20 7.2 U(VA) mL -1 / 13.2<br />

U (NBD) mL -1<br />

H 2 O 2 100 10 1<br />

2.3.4 Determination <strong>of</strong> protein concentration<br />

The determination <strong>of</strong> protein concentration was performed using the Bradford assay. This<br />

is a colorimetric protein assay based on an absorbance shift <strong>of</strong> the dye Coomassie Brilliant<br />

Blue G-250, in which under acidic conditions the red form <strong>of</strong> the dye is converted into its<br />

blue form to bind with the protein being assayed. The (bound) form <strong>of</strong> the dye has an<br />

absorption maximum historically held to be at 595 nm. The cationic (unbound) forms are<br />

green or red <strong>and</strong> absorb at 450 nm.<br />

For the calibration curve, st<strong>and</strong>ard protein (Albumin, Roth Chemie GmbH, Karlsruhe,<br />

Germany) was prepared in following concentrations: 5, 10, 20, 30, 40, 50, 60, 80, 100 µg<br />

mL -1 . 5 mL <strong>of</strong> the original dye solution Roti ® -Nanoquant (Roth Chemie GmbH, Karlsruhe,<br />

Germany) was diluted in 20 mL <strong>of</strong> distilled water. A Greiner 96-well Flat Transparent<br />

Plate was filled as follows: wells A1, A2, <strong>and</strong> A3 with 250 µL <strong>of</strong> water (blank), lines B to<br />

H (1, 2, <strong>and</strong> 3) <strong>and</strong> lines A <strong>and</strong> B (4, 5, <strong>and</strong> 6) each with 50 µL <strong>of</strong> the calibration stock<br />

45


MATERIALS AND METHODS<br />

solutions. All calibration wells were filled up with 200 µL <strong>of</strong> the dye solution. 50 µL <strong>of</strong> the<br />

protein samples were pipetted in one <strong>of</strong> the free wells <strong>and</strong> filled up with 200 µL <strong>of</strong> the dye<br />

solution. The color was visually checked <strong>and</strong> compared with a calibration row. If the color<br />

fitted into the calibration row, the sample was pipetted into three following wells (if not,<br />

the sample was first diluted) <strong>and</strong> filled up with 200 µL dye solution. The plate was<br />

measured with a multimode reader Infinite M200, TECAN® (Tecan Group Ltd.,<br />

Männedorf, Germany) with the following method: shaking for 5 seconds at 44.3 rpm,<br />

measurement at 590 nm, followed <strong>by</strong> measurement at 450 nm. The results were read as the<br />

quotient 590/450 as mean concentration in µg mL -1 .<br />

2.3.5 Difference spectra<br />

Difference spectra were recorded in quartz cuvettes (light path 10 mm) in the scanning<br />

range from 200 to 800 nm at room temperature using a Cary 50 spectrophotometer<br />

(Varian, Darmstadt, Germany). The mixture contained 50 mM potassium phosphate buffer<br />

(pH 7), 40 U mL -1 (13.75 µM) AaeAPO <strong>and</strong> 0.625 mM to 4.6 mM <strong>of</strong> the lig<strong>and</strong>. The<br />

substrate lig<strong>and</strong>s were phenol (as reference (Kinne 2010)), dicl<strong>of</strong>enac, propranolol <strong>and</strong><br />

dextromethorphan. The instrument baseline was set to zero with the reference cuvette in<br />

the sample beam. The reference cuvette contained 50 mM potassium buffer (pH 7) <strong>and</strong> 40<br />

U mL -1 (13.75 µM) <strong>of</strong> AaeAPO.<br />

2.4 St<strong>and</strong>ard reaction conditions for peroxygenase conversions<br />

Typical reaction mixtures (0.2-1.0 mL) contained purified peroxygenase (1.0-2.0 U mL -1 =<br />

0.04-0.08 µM), substrate to be oxidized (0.5-2.0 mM), potassium phosphate buffer (50<br />

mM, pH 7.0), <strong>and</strong> ascorbic acid (4.0-6.0 mM, to inhibit further oxidation <strong>of</strong> any phenolic<br />

products that were released (Kinne et al. 2009a, Kinne et al. 2008). The reactions were<br />

started <strong>by</strong> the addition <strong>of</strong> limiting H 2 O 2 (2.0-4.0 mM) <strong>and</strong> stirred at room temperature.<br />

They were stopped <strong>by</strong> addition <strong>of</strong> sodium azide (1 mM) or trichloroacetic acid (TCA) (5%)<br />

after 3 min. In some cases, H 2 O 2 was added continuously with a syringe pump <strong>and</strong> the<br />

reactions were stopped after 4-6 h, when chromatographic analyses showed that product<br />

formation was complete.<br />

Aliphatic aldehydes were analyzed as their 2,4-dinitrophenylhydrazones after addition <strong>of</strong><br />

0.2 volume <strong>of</strong> 0.1% 2,4-dinitrophenylhydrazine solution in 0.6 N HCl to each reaction<br />

mixture.<br />

46


MATERIALS AND METHODS<br />

2.5 Product identification<br />

Reaction products <strong>of</strong> interest were detected, generally with baseline resolution, <strong>by</strong> high<br />

performance liquid chromatography (HPLC) using an Agilent Series 1200 instrument<br />

equipped with a diode array detector (DAD) <strong>and</strong> an electrospray ionization mass<br />

spectrometer (MS) (Agilent Technologies Deutschl<strong>and</strong> GmbH, Böblingen, Germany).<br />

Formic acid was analyzed <strong>by</strong> ion exchange chromatography (IC) using a Dionex Series<br />

ICS 1100 system (Thermo Fisher GmbH, Idstein, Germany).<br />

2.5.1 HPLC method I<br />

For the experiments with acetaminophen, acetanilide <strong>and</strong> dicl<strong>of</strong>enac, a Synergi 4µ Fusion<br />

reversed-phase column (4.6 mm diameter <strong>by</strong> 150 mm length, 4 µm particle size, 80Ǻ,<br />

Phenomenex, Aschaffenburg, Germany) was used. The column was run at 40°C <strong>and</strong> a flow<br />

rate <strong>of</strong> 1 mL min -1 with a mixture <strong>of</strong> aqueous phosphoric acid solution (15 mM, pH 3) <strong>and</strong><br />

acetonitrile, 95:5, for 5 min, followed <strong>by</strong> a 20-min linear gradient to 100% acetonitrile.<br />

2.5.2 UHPLC method<br />

For the experiments with low concentrated propranolol, a Kinetex 2.6 µ C18 column (2.1<br />

mm diameter <strong>by</strong> 150 mm length, 2.6 µm particle size, 100 Ǻ, Phenomenex, Aschaffenburg,<br />

Germany) was used. The column was eluted at 0.3 mL min -1 <strong>and</strong> 40°C with aqueous<br />

0.01% vol/vol ammonium formate (pH 3.5)/acetonitrile, 90:10, followed <strong>by</strong> a 3-min linear<br />

gradient to 70% acetonitrile.<br />

2.5.3 HPLC method for chiral separation<br />

Chiral separation was done using the HPLC apparatus above, but using a Whelk-O 5/100<br />

Kromasil column (4.6 mm diameter <strong>by</strong> 250 mm length, Regis Technologies (Morton<br />

Grove, USA). The isocratic mobile phase consisted <strong>of</strong> 80% vol/vol methanol <strong>and</strong> 20% <strong>of</strong><br />

15 mM phosphate buffer. The columns were operated at 40°C <strong>and</strong> 1 mL min -1 for 30 min.<br />

2.5.4 HPLC-MS method I<br />

Unless <strong>other</strong>wise stated, reversed-phase (RP) chromatography <strong>of</strong> reaction mixtures was<br />

performed on a Luna C18 column (2 mm diameter <strong>by</strong> 150 mm length, 5 µm particle size,<br />

100 Ǻ, Phenomenex, Aschaffenburg, Germany), which was eluted at 0.35 mL min -1 <strong>and</strong><br />

40°C with aqueous 0.01% vol/vol ammonium formate (pH 3.5)/acetonitrile, 95:5 for 5<br />

min, followed <strong>by</strong> a 25-min linear gradient to 100% acetonitrile (for mass spectrometer<br />

parameters see 2.5.7).<br />

47


2.5.5 HPLC-MS method II<br />

MATERIALS AND METHODS<br />

The derivatized products were analyzed using a Luna C18 column. The column was eluted<br />

with aqueous 0.1% vol/vol ammonium formate (pH 3.5)/acetonitrile, 70:30 for 5 min,<br />

followed <strong>by</strong> a 20-min linear gradient to 100% acetonitrile (for mass spectrometer<br />

parameters see 2.5.7)<br />

2.5.6 HPLC-MS method III<br />

For metoprolol, a Gemini-Nx 3µ 110A C18 reversed-phase column (2 mm diameter <strong>by</strong><br />

150 mm length, 3 µm particle size, Phenomenex) was used. The column was eluted at<br />

45°C <strong>and</strong> a flow rate <strong>of</strong> 0.3 mL min -1 with a mixture <strong>of</strong> aqueous 0.2% vol/vol ammonium<br />

(pH 10.5) <strong>and</strong> acetonitrile, 95:5, for 1 min, followed <strong>by</strong> a 20-min linear gradient to 80%<br />

acetonitrile, followed <strong>by</strong> 21-min linear gradient to 100% acetonitrile (for mass<br />

spectrometer parameters see 2.5.7).<br />

2.5.7 Mass spectrometry for HPLC-MS methods<br />

Mass spectrometric determinations were made in the positive or negative ESI mode<br />

(electrospray ionization) in a mass range from 70 to 500, at a step size <strong>of</strong> 0.1, a drying gas<br />

temperature <strong>of</strong> 350°C, a capillary voltage <strong>of</strong> 4,000 V for the positive mode <strong>and</strong> 5,500 V for<br />

the negative mode. Reaction products were identified relative to authentic st<strong>and</strong>ards, based<br />

on their retention times, UV absorption spectra, <strong>and</strong> mass spectral [M + H] + or [M - H] -<br />

ions.<br />

2.5.8 IC method for organic acids<br />

Formic acid was analyzed <strong>by</strong> ion exchange chromatography (IC) on a Dionex Series ICS<br />

1100 HPLC system fitted with a IonPac® ICE-AS1 column (4 mm diameter <strong>by</strong> 250 mm<br />

length, Dionex, Germany). The column was eluted at 30°C <strong>and</strong> 1.6 mL min -1 with an<br />

aqueous perfluorobutyric acid solution (1 mM) for 20 min. The product was identified<br />

relative to an authentic st<strong>and</strong>ard <strong>by</strong> its retention time.<br />

2.6 Stoichiometrical analyses<br />

2.6.1 Sildenafil<br />

Stoichiometrical experiments on sildenafil N-demethylation were conducted in stirred<br />

reactions (0.20 mL) that contained 2 U mL -1 (0.144 µM) peroxygenase, potassium<br />

phosphate buffer (50 mM, pH 7.0) <strong>and</strong> the substrate (0.25 mM). The reactions were<br />

initiated <strong>by</strong> 0.0023-0.037 mM H 2 O 2. The resulting product (N-desmetyl sildenafil) was<br />

48


MATERIALS AND METHODS<br />

quantified <strong>by</strong> HPLC-MS as described above (HPLC-MS method I, 2.5.4) using the<br />

external st<strong>and</strong>ard curve prepared with the authentic st<strong>and</strong>ard.<br />

2.6.2 Safrole<br />

The stoichiometrical experiment on safrole demethylenation occurred in a stirred reaction<br />

mixture (0.20 mL) that contained 2 U mL -1 (0.144 µM) <strong>of</strong> peroxygenase, potassium<br />

phosphate buffer (50 mM, pH 7.0), acetonitrile (10% vol/vol), ascorbic acid (4.0 mM) <strong>and</strong><br />

the substrate (0.25 mM). The reactions were initiated <strong>by</strong> 0.0125-0.200 mM H 2 O 2 . The<br />

product (formic acid) was quantified <strong>by</strong> IC as described above, using an external st<strong>and</strong>ard<br />

curve. The st<strong>and</strong>ard curve had linear regression values <strong>of</strong> R 2 >0.99.<br />

2.7 Enzyme kinetics<br />

2.7.1 Pharmaceuticals<br />

The kinetics <strong>of</strong> propranolol hydroxylation as well as metoprolol <strong>and</strong> phenacetin O-<br />

dealkylation were analyzed in stirred microscale reactions (0.10 ml, 23°C) that contained<br />

0.40 µM peroxygenase (AaeAPO), potassium phosphate buffer (50 mM, pH 7.0), ascorbic<br />

acid (4 mM) <strong>and</strong> 0.010-5.000 mM substrate. The reactions were initiated with 2.0 mM<br />

H 2 O 2 . Reactions with propranolol or phenacetin were stopped with 0.010 mL <strong>of</strong> a 50%<br />

TCA solution after 5 seconds, <strong>and</strong> reactions with metoprolol were stopped with 0.2 volume<br />

<strong>of</strong> 0.1% 2,4-dinitrophenylhydrazine solution in 0.6 N HCl after 10 seconds. The resulting<br />

products were quantified <strong>by</strong> HPLC as described above using the external st<strong>and</strong>ard curves<br />

prepared with authentic st<strong>and</strong>ards.<br />

The kinetics <strong>of</strong> acetanilide hydroxylation were analyzed in stirred reactions (0.5 mL, 23<br />

°C) that contained 0.18 µM AaeAPO, potassium phosphate buffer (50 mM, pH 7.0),<br />

ascorbic acid (4 mM) <strong>and</strong> 0.010-5.000 mM substrate. The reactions were initiated <strong>by</strong> 2.0<br />

mM H 2 O 2 <strong>and</strong> stopped <strong>by</strong> 0.050 mL <strong>of</strong> a sodium azide solution (1 mM) after 5 seconds.<br />

The initial velocity <strong>of</strong> acetaminophen formation was then determined from the increase in<br />

absorbance at 290 nm (ε = 1,100 M -1 cm -1 ) using a Cary 50 UV/visible spectrophotometer.<br />

The kinetic parameters were determined <strong>by</strong> nonlinear regression using the Lineweaver-<br />

Burk model in the ANEMONA program (Hern<strong>and</strong>ez <strong>and</strong> Ruiz 1998).<br />

2.7.2 Nitrobenzodioxole<br />

The kinetics <strong>of</strong> 5-nitrobenzo-1,3-dioxole demethylenation were analyzed in stirred reaction<br />

mixtures (0.60 mL, 23°C ) that contained 0.036 µM (0.5 U mL -1 ) <strong>of</strong> AaeAPO, potassium<br />

phosphate buffer (50 mM, pH 7.0), <strong>and</strong> 0.125-1.000 mM <strong>of</strong> the substrate. The reactions<br />

49


MATERIALS AND METHODS<br />

were initiated with 0.063-0.100 mM H 2 O 2 , <strong>and</strong> the initial velocity <strong>of</strong> 4-nitrocatechol<br />

formation was measured <strong>by</strong> the increase in absorbance at 425 nm (ε = 9700 M -1 cm -1 )<br />

using a Cary 50 UV-Vis spectrophotometer. Three kinetic data sets were obtained for each<br />

pair <strong>of</strong> substrate concentrations. Kinetic parameters were determined <strong>by</strong> nonlinear<br />

regression using the ping-pong model in the ANEMONA program (Hern<strong>and</strong>ez <strong>and</strong> Ruiz<br />

1998).<br />

2.8 Experiments with 18 O-isotopes<br />

2.8.1 Pharmaceuticals<br />

The reaction mixtures (0.20 mL, stirred at room temperature) contained 2 U mL -1 (0.08<br />

µM) <strong>of</strong> AaeAPO, potassium phosphate buffer (50 mM, pH 7.0), <strong>and</strong> 0.5 mM substrate<br />

(tolbutamide, acetanilide, or carbamazepine). Reactions with tolbutamide or acetanilide<br />

were initiated <strong>by</strong> a single addition <strong>of</strong> 2.0 mM H 18 2 O 2 (final concentration). For reactions<br />

with carbamazepine, the same quantity <strong>of</strong> H 18 2 O 2 was added continuously with a syringe<br />

pump over 6 hours. A portion <strong>of</strong> each completed reaction was then analyzed <strong>by</strong> HPLC-MS<br />

as described above (HPLC-MS method I, 2.5.4). For each m/z value, the average total ion<br />

count within the metabolite (4-hydroxytolbutamide, acetaminophen, 3-<br />

hydroxycarbamazepine) peak was used after background correction to generate the ion<br />

count used for mass abundance calculations.<br />

2.8.2 Nitrobenzodioxole<br />

The reaction mixtures (0.20 mL, stirred at room temperature) contained 2 U mL -1 (0.144<br />

µM) <strong>of</strong> the peroxygenase, potassium phosphate buffer (50 mM, pH 7.0), ascorbic acid (4.0<br />

mM, to inhibit further oxidation <strong>of</strong> the phenolic products that were released (Kinne et al.<br />

2009a, Kinne et al. 2008) <strong>and</strong> 0.25 mM nitrobenzodioxole. The reaction was initiated with<br />

2.0 mM H 18 2 O 2 . A portion <strong>of</strong> the mixture was analyzed <strong>by</strong> HPLC-MS as described above<br />

(HPLC-MS method I, 2.5.4). Calculation <strong>of</strong> 18 O-incorporation was performed <strong>by</strong> dividing<br />

the sum <strong>of</strong> the natural species abundance <strong>and</strong> the isotope abundance with the latter. The<br />

10% <strong>of</strong> natural abundance H 2 O 2 in the H 18 2 O 2 preparation was taken into account in these<br />

calculations.<br />

2.9 Deuterium isotope effect experiment<br />

The reaction mixture (0.2 mL) contained purified AaeAPO (2 U mL -1 , 0.08 µM),<br />

potassium phosphate buffer (50 mM, pH 7.0), phenacetin-d 1 <strong>and</strong> ascorbic acid (4.0 mM).<br />

The reaction was started <strong>by</strong> the addition <strong>of</strong> H 2 O 2 (2.0 mM) <strong>and</strong> stirred at room<br />

50


MATERIALS AND METHODS<br />

temperature. The reaction was stopped <strong>by</strong> addition <strong>of</strong> 10% sodium azide (10 mM) after 10<br />

seconds. The reaction products were analyzed as described above (HPLC-MS method I,<br />

2.5.4). For each m/z value, the average total ion count within the acetaminophen peak was<br />

used after background correction to generate the ion count used for mass abundance<br />

calculations.<br />

2.10 Experiment with different peroxides <strong>and</strong> peroxide sources<br />

The reaction mixture (0.20 mL, stirred at room temperature) contained: 2 U mL -1 (0.144<br />

µM) <strong>of</strong> AaeAPO, potassium phosphate buffer (50 mM, pH 7.0), 4 mM ascorbic acid, 0.5<br />

mM <strong>of</strong> dicl<strong>of</strong>enac <strong>and</strong> 2 mM <strong>of</strong> glucose <strong>by</strong> test with glucose oxidase. Reactions were<br />

initiated in three different ways: i) single addition, ii) multiple additions (four times in four<br />

minutes), <strong>and</strong> iii) continuous addition with a syringe pump; the following peroxides or<br />

peroxide-generating systems were tested:<br />

1. 2 mM <strong>of</strong> hydrogen peroxide (H 2 O 2 ),<br />

2. 2 mM <strong>of</strong> meta-chloroperoxybenzoic acid (mCPBA),<br />

3. glucose oxidase in excess (~20 U mL -1 , in the presence <strong>of</strong> 2 mM glucose),<br />

4. 2 mM <strong>of</strong> L-cysteine,<br />

5. 2 mM <strong>of</strong> tert-butyl hydroperoxide, <strong>and</strong><br />

6. 2 mM <strong>of</strong> Cu(II) in form <strong>of</strong> copper(II) perchlorate hexahydrate [Cu(ClO 4 ) 2 × 6<br />

H 2 O].<br />

2.11 In vivo incubation experiment<br />

Eighteen 500-mL round bottom flasks were filled with the two culture media described in<br />

Table 9 (nine flasks <strong>of</strong> each medium, 200 mL).<br />

Table 9 Composition <strong>of</strong> culture media<br />

Culture medium (each 1,800 mL)<br />

Soybean medium<br />

Czapek-Dox medium<br />

54 g (3%) soybean flour 1.8 g glucose<br />

1,800 mL distilled water 5.4 g sodium nitrate<br />

1.8 g potassium hydrogen<br />

phosphate<br />

0.9 g potassium chloride<br />

0.018 g ferrous sulfate heptahydrate<br />

0.9 g yeast extract<br />

1,800 mL distilled water<br />

51


MATERIALS AND METHODS<br />

All eighteen flasks were autoclaved <strong>and</strong> numbered. Single flasks were prepared as<br />

described in Table 10.<br />

Table 10 Preparation <strong>of</strong> different cultures. a precultures <strong>of</strong> A. aegerita were prepared on MEA<br />

plates <strong>and</strong> homogenized as described under 2.2.1.1; b instead, 9 mL distilled water was<br />

supplemented; c 70% ethanol was used for the preparation <strong>of</strong> stock solution; d stock solutions: 100<br />

mM <strong>of</strong> the substance in ethanol (70%)<br />

Flask<br />

no.<br />

Czapek-Dox medium<br />

Agrocybe<br />

aegerita<br />

homogenate a<br />

(9 mL)<br />

Culture type <strong>and</strong><br />

additional<br />

substance<br />

Flask<br />

no.<br />

Soybean medium<br />

Agrocybe<br />

aegerita<br />

homogenate a<br />

(9 mL)<br />

Culture type <strong>and</strong><br />

additional<br />

substance<br />

4 + control, 1 mL <strong>of</strong><br />

1 + control, 1 mL <strong>of</strong><br />

70% ethanol c 70% ethanol c<br />

2 - b control, 1 mL<br />

propranolol d on<br />

inoculation day<br />

3 - b control, 1 mL<br />

dicl<strong>of</strong>enac d on<br />

inoculation day<br />

7 + sample, 1 mL<br />

propranolol d on<br />

inoculation day<br />

8 + sample, 1 mL<br />

dicl<strong>of</strong>enac d on<br />

inoculation day<br />

11 + sample,1 mL<br />

propranolol d one<br />

week after<br />

inoculation<br />

12 + sample, 1 mL<br />

dicl<strong>of</strong>enac d one<br />

week after<br />

inoculation<br />

15 + sample, 1 mL<br />

propranolol d two<br />

weeks after<br />

inoculation<br />

16 + sample, 1 mL<br />

dicl<strong>of</strong>enac d two<br />

weeks after<br />

inoculation<br />

5 - b 1 ml propranolol d<br />

on inoculation day<br />

6 - b control,1 mL<br />

dicl<strong>of</strong>enac d on<br />

inoculation day<br />

9 + sample, 1 mL<br />

propranolol d on<br />

inoculation day<br />

10 + sample, 1 mL<br />

dicl<strong>of</strong>enac d on<br />

inoculation day<br />

13 + sample, 1 mL<br />

propranolol d one<br />

week after<br />

inoculation<br />

14 + sample, 1 mL<br />

dicl<strong>of</strong>enac d one<br />

week after<br />

inoculation<br />

17 + sample,1 mL<br />

propranolol d two<br />

weeks after<br />

inoculation<br />

18 + sample,1 mL<br />

dicl<strong>of</strong>enac d two<br />

weeks after<br />

inoculation<br />

The culture flasks were incubated over sixteen days <strong>and</strong> agitated on a rotary shaker (100<br />

rpm) at 24°C. Sampling (1 L) <strong>of</strong> liquid from <strong>fungal</strong> cultures <strong>and</strong> controls was done in a<br />

clean bench on the day <strong>of</strong> substrate addition <strong>and</strong> the two following days. The cultures 7, 8,<br />

9, <strong>and</strong> 10 were sampled one week after inoculation. All samples were centrifuged, 500 µL<br />

were used for enzyme activity measurements, the rest was filtrated through Acrodisc® LC<br />

52


MATERIALS AND METHODS<br />

13 Syringe filters (13 mm with 0,2 µM PVDF membrane; Pall GmbH, Distributor VWR<br />

GmbH, Darmstadt, Germany) <strong>and</strong> analyzed <strong>by</strong> the HPLC-MS method I (2.5.4).<br />

2.12 <strong>Conversion</strong> <strong>of</strong> propranolol at environmentally relevant<br />

concentration<br />

Propranolol (1.5 mL; 5 µM = 5,000 nM) was added into 1,440 mL <strong>of</strong> distilled water. Six<br />

500-mL Schott® flasks were filled each with 240 mL <strong>of</strong> this diluted solution, <strong>and</strong> 9 mL <strong>of</strong><br />

distilled water was added into three control flasks; into the <strong>other</strong> three flasks, 50 Units <strong>of</strong><br />

AaeAPO dissolved in 9 mL water was added. The six flasks were incubated at 24°C<br />

overnight under mixing <strong>and</strong> continuous dosage <strong>of</strong> hydrogen peroxide (end concentration 2<br />

mM) with a syringe pump. In the next step, remaining proteins were removed from the<br />

samples; for this, the content from all six flasks was ultrafiltered through an Amicon Cell®<br />

(with 10 kDa cut-<strong>of</strong>f membrane). The membrane was conditioned with 5 µM propranolol<br />

(to prevent/reduce absorption) <strong>and</strong> the pH was adjusted to 12 with 10 M NaOH in order to<br />

facilitate subsequent extraction. The ultrafiltered samples were transferred to separation<br />

funnels, extracted with dichloromethane <strong>by</strong> vigorous shaking for 5 minutes, <strong>and</strong> left<br />

overnight for completion <strong>of</strong> phase separation. The separated organic phase was removed<br />

<strong>and</strong> concentrated in the evaporator workstation TurboVap® (Zymark, USA) to a volume <strong>of</strong><br />

1 mL; then 2 mL <strong>of</strong> acetonitrile were added <strong>and</strong> the mixture concentrated again. In the last<br />

step, 2 mL <strong>of</strong> water was added <strong>and</strong> the mixture concentrated again. The aqueous samples<br />

were analyzed <strong>by</strong> HPLC as described above (2.5.2 <strong>and</strong> 2.5.4).<br />

2.13 Long term experiments<br />

2.8.2 Preliminary test with Vivaspin® (fed-batch method)<br />

For this experiment, Vivaspin® 2 Centrifugal Concentrators (2 mL, with 10 kDa cut-<strong>of</strong>f;<br />

VivaScience Sartorius Group, Distributor VWR GmbH, Darmstadt, Germany) were used.<br />

Two Vivaspins were filled each with 10 Units <strong>of</strong> AaeAPO. Two reaction mixtures (10 mL<br />

each) were prepared, both containing 50 mM potassium phosphate buffer (pH 7), 0.50 mM<br />

propranolol, <strong>and</strong> 2 mM <strong>of</strong> hydrogen peroxide <strong>and</strong> one <strong>of</strong> them in addition 4 mM <strong>of</strong><br />

ascorbic acid. Of both mixtures 1 mL was filled in parallel into the Vivaspins that were<br />

then centrifuged with a table-centrifuge at 10,397 x g (Allegra 64R with Beckmann-Rotor<br />

C1015; Beckmann-Coulter, Krefeld, Germany). This procedure was repeated eight times<br />

on the first day <strong>and</strong> six times afterwards with one-day breaks in between (each with fresh<br />

mixtures). The Vivaspins with the enzyme were repeatedly filled with fresh potassium<br />

53


MATERIALS AND METHODS<br />

phosphate buffer after the first day <strong>and</strong> samples were stored at 4°C. After every<br />

centrifugation step, the enzyme activity was measured with the veratryl alcohol assay<br />

(2.3.1) in the permeate, as were the product (5-hydroxypropranolol) <strong>and</strong> substrate<br />

(propranolol) concentrations <strong>by</strong> HPLC-MS (method I; 2.5.4).<br />

2.8.3 Long term experiment using a hollow fiber cartridge<br />

Four 250-mL Schott ® flasks were filled with following solutions: (i) 2.5 mM propranolol;<br />

(ii) 100mM potassium phosphate buffer, pH 7; (iii) 8 mM ascorbic acid; (iv) 6.7 mM<br />

hydrogen peroxide. The outlet-tubes <strong>of</strong> the four flasks were connected with a quaternary<br />

HPLC pump (Hewlett Packard series 1050) <strong>and</strong> the solutions pumped with a flow rate <strong>of</strong><br />

0.2 ml min -1 (25% <strong>of</strong> each solution) through a hollow fiber cartridge (UFP-5-C-MM01A,<br />

GE Healtcare Europe GmbH, Freiburg, Germany). The cartridge was loaded before with<br />

30 Units (0.354 mg; A spec =84.8) <strong>of</strong> AaeAPO. For two hours, from the very beginning,<br />

samples were collected <strong>and</strong> the propranol <strong>and</strong> 5-hydroxypropranolol concentrations<br />

measured every 15 minutes <strong>by</strong> HPLC-MS (Method I; 2.5.4). This procedure was repeated<br />

for fourteen days. Every day fresh solutions (i), (ii), (iii), <strong>and</strong> (iv) were prepared <strong>and</strong><br />

pumped through the cartridge, but the enzyme in the cartridge was the same throughout the<br />

whole experiment.<br />

54


RESULTS<br />

3 Results<br />

Previous studies have shown that <strong>fungal</strong> peroxygenases catalyze various reactions similar to<br />

P450 enzymes. Pharmaceuticals <strong>and</strong> illicit <strong>drugs</strong> are one important group <strong>of</strong> P450 substrates.<br />

Their conversion <strong>by</strong> P450s usually increases solubility <strong>and</strong> bioavailability, which is<br />

prerequisite for their detoxification or further degradation in the environment. On the <strong>other</strong><br />

h<strong>and</strong>, the drug metabolites formed are <strong>of</strong> relevance for drug development <strong>and</strong> toxicological<br />

studies. In this section, the results <strong>of</strong> comprehensive tests using peroxygenase <strong>and</strong> a<br />

representative number <strong>of</strong> pharmaceutically active substances are summarized <strong>and</strong> classified<br />

according to the reaction type. Most substances were treated with peroxygenases from three<br />

different <strong>fungal</strong> species. Thereafter, the results <strong>of</strong> kinetic experiments, stoichiometrical<br />

analyses as well as <strong>of</strong> 18 O-labeling <strong>and</strong> deuterium-effect studies are presented. The next two<br />

subchapters deal with a new enzymatic assay for the determination <strong>of</strong> peroxygenase activity<br />

<strong>and</strong> experiments with selected steroids, in the course <strong>of</strong> which some catalytic limitations were<br />

observed. The last part <strong>of</strong> the results section describes complementary experiments with<br />

propranolol at low, environmentally relevant concentrations as well as long-term <strong>and</strong> in vivo<br />

studies. Due to space limitations, additional information including exemplary chromatograms<br />

<strong>and</strong> mass spectra can be found in the appendix section.<br />

3.1 Aromatic ring hydroxylation<br />

3.1.1 Ring hydroxylation <strong>of</strong> <strong>pharmaceuticals</strong><br />

Figure 16 provides an overwiew <strong>of</strong> <strong>pharmaceuticals</strong> converted <strong>by</strong> AaeAPO <strong>and</strong> CraAPO. In<br />

many cases, ring hydroxylation was observed <strong>and</strong> sometimes the preferred reaction was<br />

catalyzed <strong>by</strong> both APOs with high regioselectivity. For example, AaeAPO preferentially<br />

oxidized propranolol (I) to 5-OH-propranolol (Kinne et al. 2009a), acetanilide (II) to<br />

acetaminophen (paracetamol), carbamazepine (III) to 3-OH-carbamazepine, <strong>and</strong> dicl<strong>of</strong>enac<br />

(IV) to 4’-OH-dicl<strong>of</strong>enac (Kinne et al. 2009a). CraAPO catalyzed the same reactions but with<br />

lower regioselectivity. Both enzymes also hydroxylated ketopr<strong>of</strong>en, sulfadiazine <strong>and</strong><br />

chlorpromazine. Tamoxifen (table 15, XXVII) was oxidized <strong>by</strong> both peroxygenases yielding<br />

low amounts <strong>of</strong> 4-OH-tamoxifen along with some <strong>other</strong> products, which we were unable to<br />

resolve sufficiently <strong>by</strong> HPLC to permit quantifications.<br />

55


RESULTS<br />

Figure 16 Structures <strong>of</strong> <strong>pharmaceuticals</strong> <strong>and</strong> their ring-hydroxylated products. Chemical structures in<br />

brackets refer to structures with uncertain positions <strong>of</strong> the introduced oxygen.<br />

The observed products were also quantified, inso far as respective analytical st<strong>and</strong>ards were<br />

available. The results <strong>of</strong> quantifications are summarized in table 11.<br />

56


Table 11 Products identified <strong>by</strong> mass spectrometry after oxidation <strong>of</strong> selected <strong>pharmaceuticals</strong> <strong>by</strong> AaeAPO <strong>and</strong> CraAPO. The following data are given:<br />

m/z values <strong>of</strong> the major observed diagnostic ion for substrates <strong>and</strong> products, consumed substrate (SC), yield (Y) <strong>and</strong> regioselectivity (RS).<br />

Substrate Reaction Product<br />

Discussed as HDM<br />

AaeAPO (%) CraAPO (%)<br />

m/z<br />

in Ref. no.<br />

SC Y RS SC Y RS<br />

I 23 11<br />

5-OH-Propranolol (Otey et al. 2006) [M+H] + 276 21 91 6 53<br />

[M+H] + 260 4-OH-Propranolol (Otey et al. 2006) [M+H] + 276 − 2 13<br />

N-desisopropylpropranolol (Otey et al. 2006) [M+H] + 218 traces 3 22<br />

II 89 20<br />

Acetaminophen (Greenberg <strong>and</strong> Lester 1946) [M-H] − 150 80 90 13 65<br />

[M-H] − 134 3-OH-Acetaminophen (Greenberg <strong>and</strong> Lester 1946) [M-H] −<br />

166 5 5 1 2<br />

III 22 15<br />

3-OH-Carbamazepine (Pearce et al. 2002) [M+H] + 253 13 61 6 40<br />

[M+H] + 237 di-OH-Carbamazepine (Pearce et al. 2002) [M+H] + 269 n.d. n.d.<br />

IV 78 15<br />

[M+H] + 296 4'-OH-Dicl<strong>of</strong>enac (Bort et al. 1999) [M+H] + 312 68 87 5 30<br />

V 30 5<br />

OH-Ketopr<strong>of</strong>en<br />

[M+H] + 255<br />

(Alkatheeri et al. 1999, Skordi<br />

et al. 2005)<br />

[M+H] + 271 n.d. n.d.<br />

di-OH-Ketopr<strong>of</strong>en n.a. [M+H] + 287 n.d. n.d.<br />

VI 94 13<br />

[M+H] + 251 OH-Sulfadiazine (Winter <strong>and</strong> Unadkat 2005) [M+H] + 267 n.d. n.d.<br />

VII 99 15<br />

OH-Chlorpromazine (Perry et al. 1964) [M+H] + 335<br />

[M+H] + 319 di-OH-Chlorpromazine n.a. [M+H] + 351 n.d<br />

n.d.<br />

N-desalkyl-OH-Chlorpromazine n.a. [M+H] + 250<br />

(n.d.) = not determined due to poor resolution or no st<strong>and</strong>ard, (−) = not detected, (n.a.) = not applicable<br />

RESULTS<br />

57


RESULTS<br />

3.1.2 Ring hydroxylation <strong>of</strong> <strong>drugs</strong> <strong>of</strong> abuse<br />

Figure 17 provides an overview <strong>of</strong> controlled <strong>and</strong> illegal <strong>drugs</strong> that were ring-hydroxylated <strong>by</strong><br />

<strong>fungal</strong> peroxygenases in the presence <strong>of</strong> H 2 O 2 . The benzodiazepine <strong>drugs</strong> VIII-X were<br />

hydroxylated <strong>by</strong> AaeAPO <strong>and</strong> CraAPO to the same products, though the yields <strong>of</strong> products<br />

were lower in case <strong>of</strong> CraAPO. MroAPO oxidized only oxazepam (X).<br />

Figure 17 Structures <strong>of</strong> controlled <strong>and</strong> illicit <strong>drugs</strong> <strong>and</strong> their hypothetical ring-hydroxylated products.<br />

Chemical structures in brackets are uncertain with respect to the position <strong>of</strong> the hydroxyl group on the<br />

aromatic ring where no analytical st<strong>and</strong>ard was available.<br />

58


RESULTS<br />

As already mentioned in material <strong>and</strong> methods section, ascorbic acid was added to the<br />

reaction mixtures, to prevent the oxidative coupling <strong>and</strong> polymerization <strong>of</strong> phenolic products,<br />

<strong>by</strong> scavenging the phenoxyl radicals produced via the peroxidative activity <strong>of</strong> AaeAPO.<br />

Methamphetamine (XI), amphetamine (XII) <strong>and</strong> LSD (XIII) were also ring-hydroxylated, but<br />

the yields were very low (


RESULTS<br />

Figure 18 Structures <strong>of</strong> <strong>pharmaceuticals</strong> <strong>and</strong> their products that were oxidized at alkyl side chains.<br />

Chemical structures in brackets are hypothetical with regard to the position <strong>of</strong> the introduced oxygen<br />

(no analytical st<strong>and</strong>ard available).<br />

Figure 19 shows the structures <strong>of</strong> the APO substrate Δ 9 -tetrahydrocannabinol (THC, XVIII)<br />

<strong>and</strong> its products proposed on the basis <strong>of</strong> mass spectral <strong>and</strong> literature data (ElSohly <strong>and</strong> Feng<br />

1998).<br />

Figure 19 Structures <strong>of</strong> Δ 9 -tetrahydrocannabinol <strong>and</strong> its proposed products.<br />

60


Table 13 Products identified <strong>by</strong> mass spectrometry after oxidation <strong>of</strong> the listed <strong>pharmaceuticals</strong> <strong>by</strong> AaeAPO <strong>and</strong> CraAPO.<br />

The following data are given: m/z values for the major observed diagnostic ions for substrates <strong>and</strong> products, consumed substrate (SC), yield (Y) <strong>and</strong> regioselectivity (RS).<br />

Substrate Reaction Product<br />

Discussed as HDM<br />

AaeAPO (%) CraAPO (%)<br />

m/z<br />

in Ref no.<br />

SC Y RS SC Y RS<br />

XIV 87 98<br />

[M+H] + 207<br />

2-OH-Ibupr<strong>of</strong>en<br />

(Hamman et al. 1997, Holmes et al.<br />

2007)<br />

[M+H] + 223 21 24 74 75<br />

1-OH-Ibupr<strong>of</strong>en (Holmes et al. 2007) [M+H] + 223 7 8 −<br />

1-oxo-Ibupr<strong>of</strong>en n.a. [M+H] + 221 traces −<br />

XV a 25 20<br />

[M-H] − 271 4-OH-Tolbutamide (Hansen <strong>and</strong> Brøsen 1999) [M-H] − 287 15 60 13 62<br />

4-keto-Tolbutamide n.a [M-H] − 285 n.q. n.q.<br />

XVI 98 96<br />

OH-Gemfibrozil (Murai et al. 2004) [M-H] − 265<br />

[M-H] − 249 di-OH-Gemfibrozil (Murai et al. 2004) [M-H] − 281<br />

keto-Gemfibrozil n.a. [M-H] − 263<br />

n.q.<br />

n.q.<br />

carboxy-Gemfibrozil (Murai et al. 2004) [M-H] − 279<br />

XVII 98 96<br />

OH-Omeprazole (H<strong>of</strong>fmann 1986) [M+H] + 362<br />

n.q.<br />

Omeprazole-sulfone (Andersson et al. 1993) [M+H] + 362<br />

[M+H] + 346 keto-Omeprazole n.a. [M+H] + 362<br />

O-desmethyl-OH-Omeprazole n.a. [M+H] + 348<br />

n.q.<br />

O-desmethyl-Omeprazole-sulfone n.a. [M+H] + 348<br />

carboxy-Omeprazole (Andersson et al. 1993) [M+H] + 376<br />

XVIII 99 99<br />

[M+H] + 315<br />

OH-THC (ElSohly <strong>and</strong> Feng 1998) [M+H] + 331<br />

di-OH-THC (ElSohly <strong>and</strong> Feng 1998) [M+H] + 347<br />

n.q. n.q.<br />

a Mass spectral data indicate the formation <strong>of</strong> corresponding carbonyls as described previously (Kinne et al. 2010); (n.q.) = not quantified, no st<strong>and</strong>ard available; (−) = not<br />

detected, (n.a.) = not applicable<br />

RESULTS<br />

61


RESULTS<br />

Table 13 summarizes the m/z values <strong>of</strong> substrates <strong>and</strong> detected products, the amounts <strong>of</strong><br />

consumed substrate as well as the product yields <strong>and</strong> regioselectivity <strong>of</strong> reactions. The<br />

conversion <strong>of</strong> all substrates into the products shown in Figures 18 <strong>and</strong> 19 (XIV-XVIII)<br />

applies both to AaeAPO <strong>and</strong> CraAPO (though the product yields were different in some<br />

cases); the quantitative results are summarized in the Table 13.<br />

3.1 O-Dealkylation<br />

1.3.1 O-Dealkylation <strong>of</strong> <strong>pharmaceuticals</strong><br />

AaeAPO <strong>and</strong> CraAPO cleaved alkyl aryl ether linkages in various <strong>drugs</strong> to yield phenolic<br />

products (Kinne et al. 2009b). Thus, naproxen (XIX) was selectively demethylated to O-<br />

desmethylnaproxen, dextromethorphan (XX) to dextrorphan, <strong>and</strong> metoprolol (XXII) to O-<br />

desmethylmetoprolol.<br />

Figure 20 Structures <strong>of</strong> O-dealkylated <strong>pharmaceuticals</strong> <strong>and</strong> their products. Chemical structures in<br />

brackets are hypothetical <strong>and</strong> were deduced from mass spectral <strong>and</strong> literature data.<br />

62


Table 14 Products identified <strong>by</strong> mass spectroscopy after oxidation <strong>of</strong> the listed <strong>pharmaceuticals</strong> <strong>by</strong> AaeAPO <strong>and</strong> CraAPO in the presence <strong>of</strong> H 2 O 2 . The m/z values for the major<br />

observed diagnostic ion <strong>of</strong> substrates <strong>and</strong> products are given in each case along with the percentages <strong>of</strong> consumed substrate (SC), the yield (Y) <strong>and</strong> the regioselectivity (RS).<br />

Substrate Reaction Product<br />

Discussed as HDM<br />

AaeAPO (%) CraAPO (%)<br />

m/z<br />

in Ref no.<br />

SC Y RS SC Y RS<br />

XIX a 60 10<br />

[M-H] − 229 O-Desmethylnaproxen (Segre 1975) [M-H] − 215 57 95 9 85<br />

XX 17 15<br />

[M+H] + 272 Dextrorphan (Jacqz-Aigrain et al. 1993) [M+H] + 258 16 95 8 53<br />

XXI 34 16<br />

Acetaminophen (Brodie <strong>and</strong> Axelrod 1949) [M+H] + 152 23 66 13 80<br />

[M+H] + 136<br />

3-OH-Acetaminophen (Brodie <strong>and</strong> Axelrod 1949) [M+H] + 168 2 5 −<br />

XXII 82 73<br />

O-Desmethylmetoprolol (H<strong>of</strong>fmann et al. 1987) [M+H] + 254 17 20 4 5<br />

[M+H] + 268<br />

α-OH-Metoprolol (H<strong>of</strong>fmann et al. 1987) [M+H] + 284 2 2 1 1<br />

XXIII b − 80<br />

[M+H] + 313 Oseltamivir carboxylate (McClellan <strong>and</strong> Perry 2001) [M+H] + 285 − 71 88<br />

RESULTS<br />

a A syringe pump was used for hydrogen peroxide supply; (−) = not detected<br />

63


RESULTS<br />

Phenacetin (XXI) was deethylated to acetaminophen. Yields <strong>and</strong> regioselectivities were<br />

generally greater with AaeAPO. The above phenolic reaction products all tended to<br />

undergo further oxidation to quinones <strong>and</strong>/or coupling products because <strong>of</strong> the APOs’<br />

peroxidase activity that generates phenoxyl radicals from phenols via one-electron<br />

oxidations (Kinne et al. 2009a, Kinne et al. 2009b, Kinne et al. 2008, Ullrich <strong>and</strong><br />

H<strong>of</strong>richter 2007). Undesired coupling reactions were partially inhibitable <strong>by</strong> addition <strong>of</strong> the<br />

radical scavenger ascorbate to the reactions. DNPH derivatization <strong>of</strong> the reaction mixtures<br />

showed that the methyl group was released as formaldehyde in each case. The structures <strong>of</strong><br />

substrates <strong>and</strong> proposed products are summarized in the figure 20. In case <strong>of</strong> naproxen<br />

(XIX) <strong>and</strong> phenacetin (XXI) O-dealkylation, AaeAPO <strong>and</strong> CraAPO showed significant<br />

differences in the amount <strong>of</strong> converted substrate <strong>and</strong> the yield <strong>of</strong> products.<br />

3.1.2 Ester cleavage<br />

Ester cleavage <strong>by</strong> peroxygenase can be regarded as a special case <strong>of</strong> O-dealkylation.<br />

CraAPO was capable <strong>of</strong> catalyzing this reaction. The enzyme selectively cleaved the<br />

antiviral drug oseltamivir (XXIII) to oseltamivir carboxylate <strong>and</strong> acetaldehyde in high<br />

yield. Interestingly, oseltamivir was not converted at all <strong>by</strong> AaeAPO.<br />

3.2 N-dealkylation<br />

3.2.1 N-dealkylation <strong>of</strong> <strong>pharmaceuticals</strong><br />

APOs catalyzed the oxidative N-dealkylation <strong>of</strong> several <strong>drugs</strong> that contained secondary or<br />

tertiary amino groups. For example, AaeAPO regioselectively N-demethylated sildenafil<br />

(XXIV) at the tertiary N <strong>of</strong> the piperazine ring to give N-desmethylsildenafil <strong>and</strong><br />

formaldehyde in high yield. In contrast to that, CraAPO was ineffective in catalyzing this<br />

oxidation (figure 21). Among the <strong>other</strong> N-dealkylations showed in figure 21 are the<br />

conversion <strong>of</strong> lidocaine (XXVI) to monoethylglycinexylidide <strong>and</strong> glycinexylidide, <strong>of</strong> 4-<br />

dimethylaminoantipyrine (XXVI) to 4-aminoantipyrine <strong>and</strong> <strong>of</strong> tamoxifen (XXVII) to N-<br />

desmethyltamoxifen.<br />

64


Table 15 Products identified <strong>by</strong> mass spectrometry after oxidation <strong>of</strong> the listed <strong>pharmaceuticals</strong> <strong>by</strong> AaeAPO <strong>and</strong> CraAPO in the presence <strong>of</strong> H 2 O 2 . The m/z value for the major<br />

observed diagnostic ion <strong>of</strong> substrates <strong>and</strong> products, consumed substrate (SC), the yield (Y) <strong>and</strong> the regioselectivity (RS) are shown in each case.<br />

Discussed as HDM<br />

AaeAPO (%) CraAPO (%)<br />

Substrate Reaction Product<br />

m/z<br />

in Ref no.<br />

SC Y RS SC Y RS<br />

XXIV 82 80<br />

[M+H] + 475 N-Desmethylsildenafil (Hyl<strong>and</strong> et al. 2001) [M+H] + 461 82 99 4 5<br />

XXV 60 45<br />

Monoethylglycinexylidide (Orl<strong>and</strong>o et al. 2004) [M+H] + 207 25 41 32 70<br />

[M+H] + 235<br />

Glycinexylidide (Orl<strong>and</strong>o et al. 2004) [M+H] + 179 18 30 5 11<br />

XXVI 68 38<br />

[M+H] + 232 4-Aminoantipyrine (Banerjee et al. 1967) [M+H] + 204 16 23 19 48<br />

XXVII 25 20<br />

[M+H] + 372 N-Desmethyltamoxifen (Jin et al. 2005) [M+H] + 358<br />

4-OH-Tamoxifen (Jin et al. 2005) [M+H] + 388<br />

n.d. n.d.<br />

Endoxifen (Jin et al. 2005) [M+H] + 374<br />

(n.d.) = not determined due to poor resolution<br />

RESULTS<br />

65


RESULTS<br />

Figure 21 Structures <strong>of</strong> <strong>pharmaceuticals</strong> that were subject to N-dealkylation. Chemical structures in<br />

brackets are hypothetical.<br />

Most <strong>of</strong> the products shown in Figure 21 could be quantified; the data are summarized in<br />

Table 15.<br />

3.2.2 N-dealkylation <strong>of</strong> <strong>drugs</strong> <strong>of</strong> abuse<br />

N-Dealkylation was also observed for controlled <strong>and</strong> illicit <strong>drugs</strong> including some narcotics.<br />

The structures <strong>of</strong> the tested substrates <strong>and</strong> proposed products are shown in Figure 22.<br />

Codeine (XXVIII) <strong>and</strong> morphine (XXIX) were converted <strong>by</strong> MroAPO. The major<br />

metabolite <strong>of</strong> codeine oxidation was the N-demethylation product norcodeine. Only traces<br />

<strong>of</strong> morphine were formed, i.e. O-demethylation was much less pronounced than N-<br />

demethylation; codeine was neither a substrate <strong>of</strong> AaeAPO nor <strong>of</strong> CraAPO. Unexpectedly<br />

morphine was not subject to oxidation <strong>by</strong> MroAPO under the same conditions. Only in the<br />

absence <strong>of</strong> ascorbic acid, a conversion <strong>of</strong> morphine was observed, which – according to<br />

mass spectroscopic data - resulted in a dimerization <strong>of</strong> the molecule. This reaction might<br />

be based on the formation <strong>of</strong> phenoxyl radicals <strong>and</strong> was also observed for AaeAPO <strong>and</strong><br />

CraAPO when ascorbate was omitted.<br />

66


RESULTS<br />

Figure 22 Structures <strong>of</strong> N-dealkylated controlled <strong>pharmaceuticals</strong> <strong>and</strong> illicit <strong>drugs</strong> <strong>and</strong> their<br />

products. Chemical structures in brackets are hypothetical.<br />

The reaction <strong>of</strong> APO with cocaine (XXX) was performed under slightly acidic conditions<br />

(pH 5) to prevent the spontaneous hydrolysis <strong>of</strong> the substrate.<br />

Table 16 Products identified <strong>by</strong> mass spectrometry after oxidation <strong>of</strong> the listed <strong>pharmaceuticals</strong> <strong>by</strong><br />

AaeAPO, CraAPO <strong>and</strong> MroAPO in the presence <strong>of</strong> H 2 O 2 . The m/z values <strong>of</strong> the major observed<br />

diagnostic ion for substrates <strong>and</strong> products (P), <strong>and</strong> the amount <strong>of</strong> substrate consumed (SC) are<br />

shown in each case; (n.d.) = not determined due to poor resolution.<br />

Substrate<br />

XXVIII<br />

[M + H] + 300<br />

XXIX<br />

[M + H] + 286<br />

XXX<br />

[M + H] + 304<br />

XXXI<br />

[M + H] + 388<br />

AaeAPO CraAPO MroAPO<br />

SC P SC P SC P<br />

(%) m/z (%) m/z (%) m/z<br />

traces n.d traces n.d. 99<br />

[M + H] + 286<br />

[M + H] + 286<br />

76<br />

[M + H] + 569<br />

[M + H] + 272<br />

58<br />

[M + H] + 569<br />

[M + H] + 272<br />

0 - traces [M + H] + 290 9<br />

42<br />

[M + H] + 360<br />

[M + H] + 332<br />

30<br />

[M + H] + 360<br />

[M + H] + 332<br />

72<br />

99<br />

[M + H] + 569<br />

[M + H] + 272<br />

[M + H] + 290<br />

[M + H] + 290<br />

[M + H] + 360<br />

[M + H] + 332<br />

[M + H] + 374<br />

67


RESULTS<br />

Benzoylecgonine <strong>and</strong> norcocaine were found to be the major metabolites formed <strong>by</strong><br />

MroAPO <strong>and</strong>, in traces, <strong>by</strong> CraAPO. Flurazepam (XXXI) conversion was catalyzed <strong>by</strong> all<br />

tested peroxygenases, <strong>and</strong> single <strong>and</strong> double N-dealkylated products were detected as<br />

products. MroAPO turned out to be the best enzyme with a substrate conversion <strong>of</strong> almost<br />

100%. The results <strong>of</strong> this study are summarized in the Table 16.<br />

3.3 O-Demethylenation<br />

AaeAPO was found to oxidize diverse benzodioxoles (XXXII-XXXVIII; table 17). The<br />

reaction products were catechol derivatives <strong>and</strong> formic acid, which were identified <strong>by</strong> LC-<br />

MS <strong>and</strong> IC, respectively.<br />

Table 17 Structures <strong>of</strong> benzodioxoles <strong>and</strong> their products identified <strong>by</strong> mass spectrometry after<br />

treatment with AaeAPO in the presence <strong>of</strong> H 2 O 2 . The m/z values <strong>of</strong> the major observed diagnostic<br />

ion <strong>of</strong> the products, consumed substrate (SC), <strong>and</strong> product yield (Y) are given in each case.<br />

Substrate<br />

SC (%) Yield (%); mass ion (m/z)<br />

Name<br />

XXXII<br />

4-Nitrobenzodioxole<br />

XXXIII<br />

4-Bromobenzodioxole<br />

XXXIV<br />

4-Chlorobenzodioxole<br />

XXXV<br />

Safrole<br />

XXXVI<br />

Myristicin<br />

XXXVII<br />

MDMA („ecstasy“)<br />

O 2 N<br />

Br<br />

Cl<br />

O<br />

O<br />

O<br />

Structure<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

NH<br />

96<br />

99<br />

99<br />

78<br />

99<br />

88<br />

Demethylenated<br />

product<br />

95<br />

[M-H] - 154<br />

13<br />

[M-H] - 188<br />

75<br />

[M-H] - 143<br />

n.d.<br />

[M-H] - 149<br />

n.d.<br />

[M-H] - 179<br />

n.d.<br />

[M+H] + 182<br />

formic<br />

acid<br />

95±5.0<br />

68±2.1<br />

94±3.2<br />

75±7.0<br />

98±1.1<br />

83<br />

O<br />

XXXVIII<br />

Pisatin<br />

O<br />

OH<br />

90<br />

n.d.<br />

[M-H] - 301<br />

50<br />

O<br />

O<br />

68


RESULTS<br />

Examples <strong>of</strong> oxidized benzodioxole derivatives are the entactogenic drug ecstasy<br />

(XXXVII, MDMA, 3,4-methylenedioxymethamphetamine), the phytoalexin pisatin<br />

(XXXVIII) from Pisum sativum as well as the nutmeg-seed ingredients safrole (XXXV)<br />

<strong>and</strong> myristicin (XXXVI). Although no authentic st<strong>and</strong>ards <strong>of</strong> all reaction products were<br />

available, the m/z values for these metabolites were the same as those expected for the<br />

respective catechols with an m/z loss <strong>of</strong> 12 relative to the mass <strong>of</strong> the substrate (table 17).<br />

Moreover, we used 18 O-labeled H 2 O 2 as co-oxidant <strong>and</strong> did not find an incorporation <strong>of</strong><br />

18 O into the catechol products. Our attempts to detect 18 O-labeled formic acid failed. The<br />

reaction proceeded rapidly with total conversions between 78 <strong>and</strong> 99% <strong>of</strong> the substrate <strong>and</strong><br />

product yields up to 98%. For example, when we incubated 250 µM <strong>of</strong> 4-<br />

nitrobenzodioxole with AaeAPO under st<strong>and</strong>ard reaction conditions <strong>and</strong> non-limiting<br />

H 2 O 2 , the total conversion <strong>of</strong> the substrate was 240 µM (SC = 96 %) <strong>and</strong> we obtained<br />

237.5 µM <strong>of</strong> 4-nitrocatechol <strong>and</strong> formaldehyde (Y=95%). Some <strong>of</strong> the products were<br />

further oxidized; thus when 4-bromo- <strong>and</strong> 4-chlorobenzodioxole were converted to 4-<br />

bromo- <strong>and</strong> 4-nitrocatechol, we noticed the formation <strong>of</strong> additional ring-hydroxylated<br />

products indicated <strong>by</strong> an m/z <strong>of</strong> +16 in the mass spectrum (Appendix, Figure 8).<br />

Formaldehyde was only released from pisatin <strong>and</strong> myrisiticin, <strong>and</strong> may originate from the<br />

APO-catalyzed O-demethylation <strong>of</strong> the respective methoxy groups (Kinne et al. 2009b).<br />

3.4 Oxidation <strong>of</strong> steroids<br />

Initial attempts to convert steroids (17α-ethinyl estradiol <strong>and</strong> Reichstein’s substance S)<br />

with AaeAPO <strong>and</strong> CraAPO did not give positive results (Poraj-Kobielska et al. 2011b).<br />

Later these experiments were repeated with MroAPO <strong>and</strong> extended to <strong>other</strong> steroids.<br />

MroAPO turned out to oxidize at least some steroids including important hormones,<br />

anabolics <strong>and</strong> <strong>pharmaceuticals</strong>. In default <strong>of</strong> appropriate metabolite st<strong>and</strong>ards, the products<br />

formed are given <strong>by</strong> their m/z values in Table 18.<br />

69


RESULTS<br />

Table 18 <strong>Conversion</strong> <strong>of</strong> steroids <strong>by</strong> MroAPO. Chemical structures <strong>of</strong> the substrates, the m/z values<br />

for the major observed diagnostic ion <strong>of</strong> substrates <strong>and</strong> products, <strong>and</strong> the percentage <strong>of</strong> consumed<br />

substrates are given.<br />

70


RESULTS<br />

3.5 Stoichiometrical analyses<br />

3.5.1 Sildenafil<br />

A quantitative analysis <strong>of</strong> AaeAPO-catalyzed sildenafil oxidation in the presence <strong>of</strong><br />

limiting H 2 O 2 revealed that one equivalent <strong>of</strong> N-desmethylsildenafil was formed per<br />

equivalent <strong>of</strong> co-oxidant supplied (Table 19).<br />

71


RESULTS<br />

Table 19 Stoichiometry <strong>of</strong> sildenafil (XXIV) oxidation <strong>by</strong> AaeAPO. The initial sildenafil<br />

concentration was 250 µM.<br />

H 2 O 2 added<br />

N-desmethylsildenafil<br />

produced<br />

Ratio<br />

N-desmethylsildenafil/H 2 O 2<br />

2.3<br />

4.6<br />

9.2<br />

18.3<br />

36.6<br />

µM<br />

2.12<br />

4.27<br />

8.75<br />

20.49<br />

32.37<br />

0.92<br />

0.93<br />

0.96<br />

1.12<br />

0.88<br />

3.5.2 Safrole<br />

A quantitative analysis <strong>of</strong> safrole cleavage in the presence <strong>of</strong> a limiting amount <strong>of</strong> the cooxidant<br />

H 2 O 2 gave an atypical result. Only half an equivalent <strong>of</strong> N-desmethylsildenafil was<br />

formed per consumed equivalent <strong>of</strong> H 2 O 2 (table 20).<br />

Table 20 Stoichiometry <strong>of</strong> safrole oxidation <strong>by</strong> A. aegerita peroxygenase. The initial safrole<br />

concentration was 250 µM.<br />

H 2 O 2 added Formic acid produced Formic acid/H 2 O 2<br />

10.5<br />

17.5<br />

22.0<br />

43.8<br />

87.5<br />

µM<br />

5.70<br />

9.02<br />

11.53<br />

21.68<br />

45.53<br />

0.54 ± 0.01<br />

0.52 ± 0.01<br />

0.52 ± 0.06<br />

0.54 ± 0.06<br />

0.52 ± 0.03<br />

3.6 18 O-Isotope incorporation<br />

Using AaeAPO <strong>and</strong> three <strong>pharmaceuticals</strong> (acetanilide, carbamazepine, tolbutamide), we<br />

investigated the source <strong>of</strong> the oxygen introduced during hydroxylation. When the oxidation <strong>of</strong><br />

acetanilide (II) was conducted with H 18 2 O 2 in place <strong>of</strong> H 2 O 2 , mass spectral analysis <strong>of</strong> the<br />

resulting acetaminophen showed that the principal [M-H] − ion had shifted from its natural<br />

abundance m/z <strong>of</strong> 150 to m/z 152. Likewise, the hydroxylation <strong>of</strong> carbamazepine into 3-<br />

hydroxycarbamazepine under these conditions resulted in a shift <strong>of</strong> the [M+H] + ion from its<br />

natural abundance m/z <strong>of</strong> 253 to m/z 255. Accordingly, hydrogen peroxide was the source <strong>of</strong><br />

the introduced oxygen in both reactions.<br />

A mass spectral experiment with tolbutamide <strong>and</strong> H 18 2 O 2 as the substrates showed that the<br />

resulting 4-hydroxytolbutamide exhibited a principal [M-H] − ion with an m/z <strong>of</strong> 289 rather<br />

than 287, again showing that the peroxide was the source <strong>of</strong> transferred oxygen (Figure 23).<br />

72


RESULTS<br />

Figure 23 HPLC elution pr<strong>of</strong>iles <strong>of</strong> the<br />

conversion <strong>of</strong> tolbutamide (XV) in the<br />

presence <strong>of</strong> ascorbic acid <strong>and</strong> AaeAPO<br />

(background pr<strong>of</strong>ile), control without<br />

enzyme (front pr<strong>of</strong>ile). Insets <strong>of</strong> mass spectra<br />

show the incorporation <strong>of</strong> 18 O from H 2 18 O 2<br />

into the alcohol group <strong>of</strong> 4-<br />

hydroxytolbutamide (XVa). Upper inset:<br />

MS <strong>of</strong> the product obtained with natural<br />

abundance H 2 O 2 . Lower inset: MS <strong>of</strong> the<br />

product obtained with 90 atom% H 2 18 O 2 . The<br />

peak XVb conformed to the mass expected<br />

for 4-ketotolbutamide, an oxidation product<br />

<strong>of</strong> 4-hydroxytolbutamide (see Table 13)<br />

(Kinne et al. 2010).<br />

3.7 Kinetics experiments<br />

3.7.1 Michaelis-Menten kinetics<br />

Using AaeAPO, apparent kinetic constants were determined for some representative<br />

<strong>pharmaceuticals</strong>. The apparent K m <strong>of</strong> the the enzyme for propranolol was 1,340 µM <strong>and</strong> the<br />

k cat was 55 s -1 . For acetanilide, the constants were 1,310 µM <strong>and</strong> 1,925 s -1 . The apparent K m <strong>of</strong><br />

AaeAPO for phenacetin was 998 µM <strong>and</strong> the apparent k cat was 33 s -1 . For metoprolol, an<br />

apparent K m <strong>of</strong> 2,330 µM <strong>and</strong> a k cat <strong>of</strong> 96 s -1 were detected. All kinetic constants <strong>of</strong> AaeAPO<br />

determined here are summarized in Table 21 <strong>and</strong> compared with data obtained with <strong>other</strong><br />

substrates <strong>of</strong> AaeAPO in previous studies.<br />

Table 21 Apparent kinetic constants <strong>of</strong> AaeAPO for selected <strong>pharmaceuticals</strong> in comparison to the<br />

data for <strong>other</strong> substrates tested in previous studies (Kluge et al. 2007, Ullrich et al. 2004).<br />

Substrate<br />

Product<br />

K m<br />

(µM)<br />

V max<br />

[µM min -1 ]<br />

k cat (s -1 )<br />

k cat /K m<br />

(M -1 s -1 )<br />

ABTS (Ullrich et al. 2004) [ABTS] +· 37 283 7.7 × 10 6<br />

Acetanilide Paracetamol 1,310 21077 1925 1.5 × 10 6<br />

(Kluge et al.<br />

Naphthalene<br />

2007) 1-Naphthol 320 166 5.2 × 10 5<br />

4-Nitro-benzodioxole 4-Nitrocatechol 1,190 457 3.8 × 10 5<br />

73


RESULTS<br />

Propranolol<br />

Metoprolol<br />

5-Hydroxypropranolol<br />

Desmethylmetoprolol<br />

1,340 3781 55 4.1 × 10 4<br />

2,330 1047 96 4.1 × 10 4<br />

(Ullrich et<br />

Veratryl alcohol<br />

al. 2004) Veratraldehyde 2,367 85 3.6 × 10 4<br />

Phenacetin Paracetamol 999 361 33 3.3 × 10 4<br />

3.7.2 Bisubstrate kinetics<br />

The demethylenation <strong>of</strong> nitrobenzodioxole was chosen as a model reaction for initial rate<br />

kinetics experiments at pH 7.0, in the course <strong>of</strong> which the concentration both <strong>of</strong> the substrate<br />

<strong>and</strong> the co-substrate was varied. In contrast to the determination <strong>of</strong> the above apparent kinetic<br />

data, steady state conditions can be assumed under these conditions. The calculation <strong>of</strong> kinetic<br />

data was done using a nonlinear regression method <strong>and</strong> their double reciprocal plots gave<br />

parallel lines (figure 24), which is consistent with an enzymatic ping-pong mechanism. The<br />

kinetic parameters were as follows: k cat <strong>of</strong> 457 s -1 <strong>and</strong> a K m for H 2 O 2 <strong>of</strong> 2,160 µM <strong>and</strong> a K m for<br />

4-nitrobenzodioxole <strong>of</strong> 1,190 µM (k cat /K m = 3.8×10 5 M -1 s -1 ).<br />

Figure 24 Double reciprocal plots <strong>of</strong> the<br />

kinetic data for nitrobenzodioxole oxidation<br />

<strong>by</strong> AaeAPO. The H 2 O 2 concentrations used<br />

were 62.5 µM (■), 75 µM (●) <strong>and</strong> 100 µM<br />

(▲). The kinetic parameters reported in the<br />

text were calculated <strong>by</strong> a nonlinear<br />

regression method (Hern<strong>and</strong>ez <strong>and</strong> Ruiz<br />

1998)<br />

74


RESULTS<br />

3.8 Deuterium isotope effect experiments<br />

Phenacetin (XXI) was found to be deethylated <strong>by</strong> AaeAPO to give acetaminophen <strong>and</strong><br />

acetaldehyde (table 14, figure 20). Since phenacetin has a symmetrical site at its α-carbon, it is<br />

a suitable substrate to determine whether a catalyzed etherolytic reaction exhibits an<br />

intramolecular deuterium isotope effect. HPLC-MS analysis <strong>of</strong> DNPH-derivatized reaction<br />

mixtures showed that the AaeAPO-catalyzed cleavage <strong>of</strong> N-(4-[1- 2 H]ethoxyphenyl)acetamide<br />

(phenacetin-d 1 ) resulted in a preponderance <strong>of</strong> [ 2 H]acetaldehyde 2,4-dinitrophenylhydrazone<br />

(m/z 224, [M - H] - ) over natural abundance acetaldehyde 2,4-dinitrophenylhydrazone (m/z<br />

223, [M - H] - ) (Figure 25). The observed mean intramolecular deuterium isotope effect<br />

[(k H /k D )obs] from three experiments was 3.1 ± 0.2.<br />

Figure 25 Oxidative cleavage <strong>of</strong> N-(4-[1-<br />

2 H]ethoxyphenyl)acetamide (phenacetin-d 1 ) <strong>by</strong><br />

AaeAPO. The mass spectrum <strong>of</strong> the reaction mixture<br />

containing the products acetaldehyde 2,4-<br />

dinitrophenylhydrazone ([M - H] – , 123) <strong>and</strong> [ 2 H]<br />

acetaldehyde 2,4-dinitrophenylhydrazone ([M - H] – ,<br />

126) obtained from the oxidation <strong>of</strong> phenacetin-d 1 is<br />

shown. The mass spectrum is one <strong>of</strong> three used to<br />

calculate the observed mean intramolecular isotope<br />

effect.<br />

3.9 Spectrophotometric nitrobenzodioxole assay<br />

Peroxygenases from A. aegerita, C. radians, Coprinopsis verticillata <strong>and</strong> M. rotula as well as<br />

recombinant peroxygenase <strong>of</strong> Coprinopsis cinerea were dissolved together with<br />

nitrobenzodioxole (NBD) in potassium phosphate buffer containing acetonitrile (10% vol/vol)<br />

<strong>and</strong>, if the reaction products were to be quantified <strong>by</strong> HPLC, ascorbic acid. After adding the<br />

co-substrate H 2 O 2 , the reaction mixture immediately changed its color to bright yellow (figure<br />

26, inset A, Figure 27).<br />

The metabolite causing color change was identified as 4-nitrocatechol <strong>by</strong> HPLC-MS relative<br />

to an authentic st<strong>and</strong>ard with complete baseline resolution (compare also Table 17). The<br />

reaction proceeded selectively with almost total conversions up to 99% <strong>of</strong> NBD. The<br />

concentration-dependent change <strong>of</strong> absorbance in the reaction mixture (caused <strong>by</strong> the<br />

formation <strong>of</strong> 4-nitrocatechol) at wavelengths in the visible region <strong>and</strong> at different stages <strong>of</strong> the<br />

reaction (figure 26 A) was used to develop a direct spectrophotometric assay. A<br />

spectrophotometer was applied to follow this rapid reaction over 15-30 seconds, which<br />

corresponded with the linear initial slope (Figure 26 B). To prove the applicability <strong>of</strong> the new<br />

enzyme assay, we followed the secretion <strong>of</strong> A. aegerita aromatic peroxygenase (AaeAPO) in<br />

75


RESULTS<br />

liquid cultures spectrophotometrically with the NBD assay <strong>and</strong> compared the results with the<br />

classic peroxygenase assay 9 (Figure 27).<br />

Figure 26 Concentration <strong>of</strong> 4-nitrocatechol in reactions followed <strong>by</strong> HPLC vs. absorbance at 425 nm<br />

measured spectrophotometrically. Parallel reactions <strong>of</strong> the AaeAPO-catalyzed conversion <strong>of</strong> NBD<br />

(0.125 mM) were stopped after different times <strong>of</strong> incubation (1: 5 s; 2: 15 s; 3: 30 s; 4: 60 s; 5: 120 s 6:<br />

240 s; 7: 360 s) using sodium azide (1 mM). Insets show the time course <strong>of</strong> 4-nitrocatechol production<br />

(A) <strong>and</strong> the spectral changes during the conversion (B).<br />

It was found that the absorbance increase <strong>of</strong> the reaction mixture at 425 nm measured<br />

spectrophotometrically correlated well (R 2 =0.99) with the concentration <strong>of</strong> produced 4-<br />

nitrocatechol detected <strong>by</strong> HPLC in the range from 10 to 100 μM (Figure 26). Based on this<br />

result <strong>and</strong> additional spectrophotometric measurements with NBD <strong>and</strong> 4-nitrocatechol<br />

st<strong>and</strong>ards, we calculated a corrected extinction coefficient 10 <strong>of</strong> 9,700 M −1 cm −1 for 4-<br />

nitrocatechol at 425 nm under the reaction conditions used.<br />

9 1 U (unit) <strong>of</strong> peroxygenase represents the oxidation <strong>of</strong> 1 µmol <strong>of</strong> 3,4-dimethoxybenzyl<br />

alcohol into 3,4-dimethoxybenzaldehyde in 1 min at 23 °C (Ullrich et al. 2004)<br />

10 The actual extinction coefficient was calculated from calibration curves spectrophotometrically recorded for 4-<br />

nitrocatechol at 425 nm (ε = 9,900 M -1 cm -1 ) <strong>and</strong> corrected <strong>by</strong> subtraction <strong>of</strong> the extinction coefficient <strong>of</strong> NBD at<br />

425 nm (ε = 200 M -1 cm -1 ).<br />

76


RESULTS<br />

Figure 27 Parallel measurements <strong>of</strong> AaeAPO activity during liquid cultivation <strong>of</strong> the fungus A.<br />

aegerita (days 9 to 23) in a soybean medium (2% w/vol) using the veratryl alcohol assay (I) (Ullrich et<br />

al. 2004) <strong>and</strong> the new NBD assay (II). The inset shows the change <strong>of</strong> color before (A; pH = 7) <strong>and</strong><br />

after the addition <strong>of</strong> sodium hydroxide (B; pH >12).<br />

The results show that the measured volume activity <strong>of</strong> the enzyme in the supernatant <strong>of</strong> the<br />

culture liquid correlates well with the color change over a broad activity range. The activity <strong>of</strong><br />

A. aegerita peroxygenase measured with NBD was app. 80% higher than that with veratryl<br />

alcohol (may be due to a higher turnover number <strong>of</strong> the enzyme for NBD). After alkalization<br />

(pH >12) <strong>of</strong> the yellow reaction mixture with 10% (vol/vol) <strong>of</strong> 1 M sodium hydroxide, its<br />

color changed to deep red (Figure 27, inset B). The corresponding UV-Vis spectrum <strong>of</strong> 4-<br />

nitrocatechol showed an absorption maximum at 514 nm (ε 514 = 11,400 M −1 cm −1 , Appendix<br />

Figure 9) (Cooper <strong>and</strong> Tulane 1936a). To confirm the specificity <strong>of</strong> the assay, we tested a<br />

number <strong>of</strong> secreted oxidoreductases, such as horseradish peroxidase, soybean peroxidase,<br />

chloroperoxidase from L. fumago <strong>and</strong> laccase from Trametes versicolor, none <strong>of</strong> which<br />

showed any activity towards NBD. On the <strong>other</strong> h<strong>and</strong>, the aromatic peroxygenases from C.<br />

radians, C. verticillata <strong>and</strong> M. rotula as well as the recombinant enzyme <strong>of</strong> C. cinerea<br />

converted NBD into 4-nitrocatechol (Appendix Figure 7).<br />

To prove the applicability <strong>of</strong> the NBD assay for environmental samples (at pH 7), we added<br />

extracts (300 µL) from forest litter, wheat straw <strong>and</strong> beech wood to the assay mixture (0.7<br />

mL) that contained a defined amount <strong>of</strong> AaeAPO (0.067 µM = 0.166 U mL -1 ) (see Chapter<br />

2.3.3.3). Relative to the control assay, 87 ± 3% <strong>of</strong> added enzyme activity was detected in the<br />

77


RESULTS<br />

presence <strong>of</strong> the forest litter extract as well as 74 ± 5% <strong>and</strong> 55 ± 8% in the presence <strong>of</strong> the<br />

straw <strong>and</strong> beech wood extracts, respectively.<br />

3.10 <strong>Conversion</strong> <strong>of</strong> a set <strong>of</strong> human P450 substrates<br />

A set <strong>of</strong> twelve common human P450 substrates (<strong>pharmaceuticals</strong>, xenobiotics, plant<br />

ingredients), that is routinely used to distinguish between different CYP types (Walsky <strong>and</strong><br />

Obach 2004), was tested with our three model peroxygenases (AaeAPO, CraAPO, MroAPO).<br />

The results are summarized in Table 22.<br />

Coumarin (LVII; specific for CYP2A6) <strong>and</strong> felodipine (LV; CYP3A) were not or hardly<br />

(


RESULTS<br />

Table 22 Results <strong>of</strong> the treatment <strong>of</strong> twelve common substrates <strong>of</strong> human cytochrome P450s (CYPs)<br />

<strong>by</strong> the <strong>fungal</strong> peroxygenases from A. aegerita, C. radians <strong>and</strong> M. rotula. P450/CYP products were<br />

adopted from following references: (Walsky <strong>and</strong> Obach 2004). 1 (+) P450 metabolites were found,<br />

(±) indication for the formation <strong>of</strong> P450 metabolites was found but st<strong>and</strong>ard was lacking, (-) no P450<br />

metabolites were found; # rapid further conversion into the corresponding ketone product.<br />

Substrate AaeAPO CraAPO MroAPO<br />

LII<br />

<strong>Conversion</strong> 6 15 16<br />

Amodiaquine (%)<br />

Cl N P450-<br />

product<br />

(CYP2C8)<br />

desethylamodiaquine<br />

desethylamodiaquine<br />

Cl<br />

N<br />

H O<br />

LIII<br />

Bupropion<br />

O<br />

NH<br />

NH<br />

APO<br />

Yield (%)<br />

desethylamodiaquine<br />

<strong>Conversion</strong><br />

(%)<br />

P450-<br />

product<br />

(CYP2B6)<br />

+<br />

4<br />

+<br />

10<br />

+<br />

11<br />

27 26 18<br />

Cl<br />

Cl<br />

LIV<br />

Chlorzoxazone<br />

H<br />

N<br />

O<br />

O<br />

O<br />

LV<br />

Felodipine<br />

Cl<br />

Cl<br />

O<br />

N<br />

H<br />

LVI<br />

Midazolam<br />

N<br />

N<br />

N<br />

F<br />

O<br />

O<br />

APO ±<br />

(3 hydroxylated<br />

products)<br />

<strong>Conversion</strong><br />

hydroxybupropion<br />

hydroxybupropion<br />

hydroxybupropion<br />

(%)<br />

P450-<br />

product<br />

(CYP2E1)<br />

±<br />

(2 hydroxylated<br />

products)<br />

58 38 8<br />

6-hydroxy<br />

chlorzoxazone<br />

APO ±<br />

(1 hydroxylated<br />

product)<br />

<strong>Conversion</strong><br />

(%)<br />

P450-<br />

product<br />

(CYP3A)<br />

6-hydroxy<br />

chlorzoxazone<br />

±<br />

(1 hydroxylated<br />

product)<br />

-<br />

6-hydroxy<br />

chlorzoxazone<br />

±<br />

(1 hydroxylated<br />

product)<br />


RESULTS<br />

Coumarin (%)<br />

O O<br />

P450-<br />

product<br />

(CYP2A6)<br />

7-hydroxycoumarin<br />

7-hydroxycoumarin<br />

7-hydroxycoumarin<br />

APO - - -<br />

LVIII <strong>Conversion</strong> 15 no data 11<br />

S-Mephenytoin (%)<br />

H<br />

P450- 4-hydroxy-Smephenytoimephenytoin<br />

no data 4-hydroxy-S-<br />

N O<br />

product<br />

N (CYP2C19)<br />

APO + (traces) + (traces)<br />

XXI <strong>Conversion</strong> 34 16 traces<br />

Phenacetin<br />

(%)<br />

O<br />

P450- acetaminophen acetaminophen acetaminophen<br />

NH<br />

product<br />

(CYP1A2)<br />

O<br />

APO/<br />

+<br />

+<br />

-<br />

Yield(%) 23<br />

13<br />

0<br />

IV<br />

<strong>Conversion</strong> 78 15 68<br />

Dicl<strong>of</strong>enac<br />

(%)<br />

Cl<br />

P450-<br />

product<br />

4’-hydroxydicl<strong>of</strong>enac<br />

4’-hydroxydicl<strong>of</strong>enac<br />

4’-hydroxydicl<strong>of</strong>enac<br />

NH<br />

(CYP2C9)<br />

Cl<br />

O<br />

OH<br />

APO/<br />

Yield(%)<br />

4-hydroxytolbutamide<br />

XV<br />

Tolbutamide<br />

<strong>Conversion</strong><br />

(%)<br />

P450-<br />

product<br />

(CYP2C9)<br />

O HN<br />

APO/<br />

S NH O yield(%)<br />

O<br />

XX<br />

<strong>Conversion</strong><br />

Dextromethorphan (%)<br />

N H<br />

P450-<br />

product<br />

(CYP2D6)<br />

APO/<br />

yield(%)<br />

+<br />

68<br />

+<br />

5<br />

+<br />

4<br />

25 20 29<br />

4-hydroxytolbutamide<br />

4-hydroxytolbutamide<br />

+<br />

+<br />

+<br />

15 # 13 #


RESULTS<br />

3.11 Difference spectroscopy<br />

Heme containing enzymes can be characterized using a number <strong>of</strong> spectroscopic methods.<br />

Thus it is known that the addition <strong>of</strong> various substrates or inhibitors can cause characteristic<br />

changes in the optical absorption spectrum <strong>of</strong> P450 enzymes. The spectral changes reflect the<br />

binding <strong>of</strong> these compounds near the active site, which causes changes in the heme<br />

environment <strong>and</strong> influences the electron density (Lewis 2001).<br />

Lewis described three characteristic types <strong>of</strong> substrate binding spectra for P450s, namely type<br />

I (e.g. with phenol or propranolol), type II (e.g. with cyanide or pyridine), <strong>and</strong> the reverse type<br />

I (sometimes referred as modified type II, e.g. with ethanol or caffeine) (Lewis 2001,<br />

Schenkman et al. 1972, Schenkman et al. 1981).<br />

⎯ Type I is characterized <strong>by</strong> a decrease in intensity <strong>of</strong> the Soret absorption peak around<br />

418 nm, coupled with a concominant increase in the b<strong>and</strong> intensity around 390 nm.<br />

Consequently, a type I spectrum shows that the binding <strong>of</strong> substrate within the P450<br />

heme locus brings about a shift in the iron(III) spin equilibrium from low-spin to highspin<br />

(Lewis 1996).<br />

⎯ Type II is typical for compounds acting as P450 inhibitors via ligation <strong>of</strong> the heme<br />

iron (Lewis 1996). The UV spectrum is characterized <strong>by</strong> decrease in absorption around<br />

390 nm with concominant increase around 430 nm. Consequently, there is a shift in<br />

spin-state equilibrium from high-spin to low-spin iron.<br />

⎯ Reverse type I is a "mirror-image" change to type I, where increase <strong>of</strong> absorption at<br />

420 nm is coupled with a decrease in the 390 nm peak (Lewis 2001).<br />

The results in Figure 28 <strong>and</strong> Table 23 show that the substrates propranolol, dicl<strong>of</strong>enac <strong>and</strong><br />

dextromethorphan (Kinne et al. 2009a, Poraj-Kobielska et al. 2011b) as well as phenol<br />

(Ullrich et al. 2004) caused a type I spectral change with the formation <strong>of</strong> maxima between<br />

385 <strong>and</strong> 398 nm <strong>and</strong> minima around 420 nm.<br />

Table 23 Difference absorption spectra <strong>of</strong> AaeAPO obtained <strong>by</strong> the addition <strong>of</strong> three common<br />

<strong>pharmaceuticals</strong> <strong>and</strong> phenol as a reference.<br />

Lig<strong>and</strong> λ max (nm) λ min (nm) Spectral type<br />

Phenol 1 388 421 I<br />

Dextromethorphan 389 422 I<br />

Propranolol 386 422 I<br />

Dicl<strong>of</strong>enac 386 417 I<br />

1 (Kinne 2010)<br />

81


RESULTS<br />

Figure 28 Difference absorption spectra <strong>of</strong> AaeAPO obtained <strong>by</strong> the addition <strong>of</strong> phenol (A),<br />

dextromethorphan (B), propranolol (C) und dicl<strong>of</strong>enac (D).<br />

3.12 Different peroxides <strong>and</strong> peroxide-generating systems<br />

The conversion <strong>of</strong> dicl<strong>of</strong>enac into 4'-hydroxydicl<strong>of</strong>enac <strong>by</strong> AaeAPO was used to test the<br />

suitability <strong>of</strong> different peroxides <strong>and</strong> peroxide-generation systems. The best result was<br />

obtained with hydrogen peroxide (H 2 O 2 ) added in a few small doses, the lowest conversion<br />

was observed with the system <strong>of</strong> L-cysteine. The organic peroxides mCPBA <strong>and</strong> tert-butyl<br />

82


RESULTS<br />

hydroperoxide were less efficient than H 2 O 2 but nevertheless enabled AaeAPO to catalyze the<br />

substantial oxidation <strong>of</strong> dicl<strong>of</strong>enac. The results <strong>of</strong> this study are summarized in Table 24.<br />

Table 24 Comparison <strong>of</strong> different peroxides <strong>and</strong> peroxide-generating systems. The experiments were<br />

performed with dicl<strong>of</strong>enac (0.5 mM), AaeAPO (2 U mL -1 ) in the presence <strong>of</strong> ascorbic acid <strong>and</strong> 2<br />

mM <strong>of</strong> each peroxide. Details are described in Chapter 2.10.<br />

Peroxide/<br />

peroxide generating system<br />

Yield <strong>of</strong> 4'-OH-<br />

Dicl<strong>of</strong>enac (%)<br />

H 2 O 2 (one time addition) 42<br />

H 2 O 2 (few doses) 45<br />

H 2 O 2 (syringe pump) 34<br />

mCPBA 25<br />

mCPBA (few doses) 32<br />

mCPBA (syringe pump) 27<br />

Glucose/glucose oxidase 36<br />

L-cysteine 6<br />

tert-butyl hydroperoxide 13<br />

Copper(II) perchlorate hexahydrate/ascorbic acid 12<br />

3.13 In vivo experiments<br />

Growing liquid cultures <strong>of</strong> A. aegerita were supplemented with propranolol or dicl<strong>of</strong>enac<br />

(500 µM; 500 ppm); controls lacked either the <strong>pharmaceuticals</strong> or the fungus. Pictures <strong>of</strong> the<br />

cultures are shown in Figure 29.<br />

A<br />

B<br />

Figure 29 Liquid cultures <strong>of</strong> A. aegerita: on day 8 (A): soybean medium, one day after addition <strong>of</strong><br />

dicl<strong>of</strong>enac (left) <strong>and</strong> control without dicl<strong>of</strong>enac (right); (B): Czapek-Dox medium, one day after<br />

addition <strong>of</strong> dicl<strong>of</strong>enac (right) <strong>and</strong> without dicl<strong>of</strong>enac (left).<br />

Figure 30 summarizes the results <strong>of</strong> this in vivo experiment, in the course <strong>of</strong> which both<br />

propranolol <strong>and</strong> dicl<strong>of</strong>enac disappeared. This effect was more pronounced in soybean cultures<br />

that produce much more peroxygenase than Czapek-Dox cultures do (Ullrich et al. 2004), but<br />

where also adsorption processes may significantly contribute to the substrate removal.<br />

83


84<br />

RESULTS<br />

Figure 31 HPLC elution pr<strong>of</strong>iles <strong>of</strong> samples from <strong>fungal</strong> cultures containing dicl<strong>of</strong>enac (IV) or propranolol (I) in soybean medium. (A) Culture <strong>of</strong> A. aegerita with dicl<strong>of</strong>enac on<br />

day 14 when the pharmaceutical was added <strong>and</strong> on day 16, compared to a control without fungus; (B) Fungal culture with propranolol on day 7 (when substrate was added) <strong>and</strong><br />

the day after, compared to a control without fungus. (II) <strong>and</strong> (III) refer to medium components. The insets show the DAD-spectra <strong>of</strong>: (I) propranolol with structure; (II) <strong>and</strong> (III)<br />

unidentified soybean medium components; (IV) dicl<strong>of</strong>enac with structure.


RESULTS<br />

Figure 30 Propranolol <strong>and</strong> dicl<strong>of</strong>enac concentration in in-vivo experiments with liquid cultures <strong>of</strong> A.<br />

aegerita: soybean medium (red), Czapek-Dox (green).<br />

No indications for the formation <strong>of</strong> the typical peroxygenase metabolites were found.<br />

However, both substrates disappeared rapidly in the <strong>fungal</strong> culures, particularly in those<br />

containing soybean flour (where the enzyme activity was near its maximum around day 14;<br />

Figure 30). Interestingly the detectable concentration <strong>of</strong> both substrates in the soybean<br />

cultures decreased <strong>by</strong> more than 50% immediately after their addition (also in the controls<br />

without fungus). This was probably the result <strong>of</strong> substrate absorption to the soybean particles<br />

(figure 29 A). Figure 31 shows, as an example, the changes in the HPLC elution pr<strong>of</strong>iles<br />

during the incubation <strong>of</strong> propranolol <strong>and</strong> dicl<strong>of</strong>enac with A. aegerita in the soybean medium.<br />

3.14 Tests with environmentally relevant concentrations <strong>of</strong><br />

propranolol<br />

The concentration <strong>of</strong> <strong>pharmaceuticals</strong> in the environment (e.g. waste water, activated sludge)<br />

is usually rather low (i.e. between 0.1-1000 ppb). Thus the testing <strong>of</strong> environmetally relevant<br />

propranolol concentrations required the optimization <strong>of</strong> the analytical HPLC method <strong>and</strong><br />

sample preparation (2.5.2 <strong>and</strong> 2.12.). As a result, it was possible to identify <strong>and</strong> quantify<br />

85


propranolol in the nM range (M W : 259 g mol -1 , 1 nM = 0.259 µg L -1 ). The outcome <strong>of</strong> this<br />

experiment was that about 50% <strong>of</strong> the substrate propranolol at a concentration as low as 5 nM<br />

(1.3 ppb) could be removed <strong>by</strong> AaeAPO (figure 32).<br />

Figure 32 HPLC elution<br />

pr<strong>of</strong>iles showing the<br />

disappearance <strong>of</strong><br />

propranolol at an<br />

environmentally relevant<br />

concentration (5 nM = 1.3<br />

ppb = 1.3 µg L -1 ) after<br />

treatment with 50 units <strong>of</strong><br />

AaeAPO for 20 hours.<br />

3.15 Long-term experiments<br />

3.15.1 Pre-study with Vivaspin® (fed-batch method)<br />

A Vivaspin® centrifugal concentrator (molecular mass cut-<strong>of</strong>f 10 kDa) was used to determine<br />

whether AaeAPO is stable enough to be repeatedly used in enzymatic conversions. Since the<br />

concentrator holds back proteins, the enzyme, can in principle be used several times (Figure<br />

33). Experiments with propranolol as the substrate demonstrated that this approach is feasible.<br />

Surprisingly, the experiment without ascorbic acid gave more<br />

product (5-OH-propranolol) than the experiment in its presence<br />

(Figure 34), indicating – despite the reduction <strong>of</strong> phenoxyl radicals<br />

that ascorbate provides – a general lowering <strong>of</strong> the overall<br />

efficiency <strong>of</strong> the peroxygenase system <strong>by</strong> the radical scavenger.<br />

A<br />

B<br />

Figure 33 Vivaspin tubes after treatment <strong>of</strong> propranolol with A. aegerita<br />

peroxygenase (A) in the absence <strong>of</strong> ascorbic acid <strong>and</strong> (B) in its presence.<br />

86


RESULTS<br />

In the absence <strong>of</strong> ascorbic acid, the formation <strong>of</strong> dark substances was observed in the<br />

Vivaspins®, which can be attributed to the oxidative coupling <strong>of</strong> phenoxyl radicals formed<br />

from 5-OH-propranolol <strong>by</strong> the peroxidase activity <strong>of</strong> AaeAPO. The results <strong>of</strong> this experiment<br />

are summarized in Figure 34.<br />

Figure 34 Production <strong>of</strong> 5-OH-propranolol (green columns) <strong>and</strong> presence <strong>of</strong> remaining propranolol<br />

(red columns) in Vivaspin experiments with AaeAPO. Reactions with (left) <strong>and</strong> without (right)<br />

ascorbic acid.<br />

3.15.2 Long-term experiment using a hollow fiber cartridge<br />

Figure 35 Production <strong>of</strong> 5-OH-propranolol in a long-term experiment using AaeAPO. Daily amount <strong>of</strong><br />

propranolol added (green columns); cumulative amount <strong>of</strong> formed 5-OH-propranolol (orange<br />

columns).<br />

The preliminary experiments with Vivaspin® demonstrated that AaeAPO is highly stable <strong>and</strong><br />

can be used in long-term experiments. For that, the ultrafiltration membrane <strong>of</strong> a fiber<br />

87


cartridge was loaded with AaeAPO protein <strong>and</strong> continuously used for propranolol<br />

hydroxylation over 14 days. Figure 35 illustrates the progress <strong>of</strong> this experiment. In all, 13.85<br />

mg <strong>of</strong> 5-OH-propranolol was produced from 217.56 mg <strong>of</strong> propranolol using 354 µg enzyme<br />

over a period <strong>of</strong> fourteen days.<br />

3.16 Scope <strong>of</strong> peroxygenase reactions concerning<br />

<strong>pharmaceuticals</strong> <strong>and</strong> <strong>drugs</strong><br />

Figure 36 Substrates that were not converted <strong>by</strong> <strong>fungal</strong> peroxygenases.<br />

88


RESULTS<br />

In the course <strong>of</strong> this PhD project, some limitations <strong>of</strong> <strong>fungal</strong> peroxygenases concerning the<br />

substrate spectrum have been ascertained. No conversion was observed for cocaine (XXX),<br />

codeine (XXVIII), testosterone (XLI), 17α-ethinylestradiol (XLII), Reichstein’s substance S<br />

(XLIII) <strong>and</strong> felbamate (LIX) when treated with AaeAPO or CraAPO. Furthermore, AaeAPO<br />

was not able to convert oseltamivir (XXIII). The experimental results furthermore indicate<br />

that APOs <strong>of</strong>ten fail to discriminate between chiral centers in the pharmaceutical substrates.<br />

For example, chiral HPLC separation <strong>of</strong> the two O-desmethylnaproxen entantiomers that<br />

resulted from naproxen oxidation showed that neither enantiomer predominated in the endproduct<br />

mixture. In case <strong>of</strong> MroAPO, no product formation was observed for temazepam<br />

(IX), oxazepam (X), diazepam (VIII) <strong>and</strong> amphetamine (XII). All three enzymes were unable<br />

to convert felodipine (LV) <strong>and</strong> coumarine (LVII). The structures <strong>of</strong> the non-oxidizable<br />

substrates are given in Figure 36.<br />

89


DISCUSSION<br />

4 Discussion<br />

4.1 Preparation <strong>of</strong> human drug metabolites<br />

Human drug metabolites (HDMs) are valuable chemicals needed for the development <strong>of</strong><br />

safe <strong>and</strong> effective <strong>pharmaceuticals</strong>. They are frequently used as substrates <strong>and</strong> authentic<br />

st<strong>and</strong>ards in studies <strong>of</strong> drug bioavailability, pharmacodynamics <strong>and</strong> pharmacokinetics<br />

(Baillie et al. 2002, Nelson 1982, Peters et al. 2007). In vivo, the C-H bonds <strong>of</strong><br />

<strong>pharmaceuticals</strong> are predominantly oxygenated <strong>by</strong> liver cytochrome P450-<br />

monooxygenases (P450s) to yield more polar HDMs that are excreted directly or as<br />

conjugates (Gibson <strong>and</strong> Skett 2001). This directed incorporation <strong>of</strong> an oxygen atom into a<br />

complex organic structure is one <strong>of</strong> the most challenging reactions in synthetic organic<br />

chemistry (Ullrich <strong>and</strong> H<strong>of</strong>richter 2007). Thus low yields <strong>and</strong> the need for laborious<br />

removal <strong>of</strong> <strong>by</strong>products have prevented the cost-effective preparation <strong>of</strong> HDMs <strong>by</strong> purely<br />

chemical methods (Kinne et al. 2009a). A preferable route to prepare human drug<br />

metabolites would be the use <strong>of</strong> isolated human P450s, but these complex multiprotein<br />

systems are membrane-bound, poorly stable, c<strong>of</strong>actor-dependent, <strong>and</strong> generally exhibit<br />

low reaction rates (Eiben et al. 2006, Urlacher et al. 2004). More promising results have<br />

been reported for human cell cultures (Acikgöz et al. 2009), microbial whole-cell systems<br />

that heterologouly express human P450s (Peters et al. 2007) <strong>and</strong> laboratory-evolved<br />

bacterial P450s (Damsten et al. 2008, Sawayama et al. 2009). The latter are capable <strong>of</strong><br />

catalyzing the H 2 O 2 -dependent hydroxylation <strong>of</strong> <strong>pharmaceuticals</strong> via the ”peroxide shunt”<br />

pathway (Otey et al. 2006). Recent studies, however, have demonstrated that this approach<br />

needs further optimization, in particular regarding turnover numbers <strong>and</strong> peroxide stability<br />

(Sawayama et al. 2009). Alternatively, modified hemoproteins such as microperoxidases<br />

might be used to catalyze hydroxylations <strong>by</strong> a P450-like oxygen transfer mechanism, but<br />

so far these catalysts have not reached the necessary performance <strong>and</strong> selectivity (Caputi et<br />

al. 2005, Dallacosta et al. 2003, Dorovska-Taran et al. 1998, Osman et al. 1996, Prieto et<br />

al. 2006, Veeger 2002).<br />

A few years ago, a new heme-thiolate enzyme was discovered in the wood- <strong>and</strong> litterdwelling<br />

black poplar mushroom, Agrocybe aegerita. This enzyme turned out to be a true<br />

peroxygenase (aromatic/unspecific peroxygenase, APO; EC 1.11.2.1), efficiently<br />

transferring oxygen from peroxides to various organic substrates including aromatic,<br />

heterocyclic <strong>and</strong> aliphatic compounds (H<strong>of</strong>richter et al. 2010, Ullrich <strong>and</strong> H<strong>of</strong>richter 2005,<br />

90


DISCUSSION<br />

Ullrich et al. 2004). In the following years, a few similar enzymes have been found in<br />

cultures <strong>of</strong> <strong>other</strong> agaric basidiomycetes such as Coprinellus radians, Coprinopsis<br />

verticilata (Anh et al. 2007) <strong>and</strong> Marasmius rotula (Gröbe et al. 2012) These stable,<br />

secreted enzymes are promising oxidoreductases for biotechnological applications<br />

including the synthesis <strong>of</strong> human drug metabolites. Depending on the particular substrate<br />

<strong>and</strong> the reaction conditions, APOs catalyze in addition to peroxygenations, also oneelectron<br />

oxidations (e.g. <strong>of</strong> phenols) <strong>and</strong> unspecific halogenations (bromination), <strong>and</strong> for<br />

all these reactions, they need merely a peroxide as co-substrate to function.<br />

Previous studies with diverse organic compounds showed that APOs catalyze several types<br />

<strong>of</strong> oxyfunctionalization such as epoxidation, aromatic <strong>and</strong> benzylic hydroxylation, ether<br />

cleavage (O-dealkylation), N-dealkylation, sulfoxidation <strong>and</strong> N-oxidation (Ar<strong>and</strong>a et al.<br />

2009a, Ar<strong>and</strong>a et al. 2009b, H<strong>of</strong>richter et al. 2010, Kinne et al. 2009a, Kinne et al. 2009b,<br />

Kinne et al. 2008, Kinne et al. 2010). All these reactions are also typical <strong>of</strong> P450s <strong>and</strong> part<br />

<strong>of</strong> drug transformations in the human liver (Guengerich 1987, Ortiz de Montellano 2005).<br />

Promising results were achieved in the prestudies with the <strong>pharmaceuticals</strong> propranolol<br />

<strong>and</strong> dicl<strong>of</strong>enac, which were converted regioselectively into the respective hydroxylated<br />

human drug metabolites (Kinne et al. 2009a). All <strong>of</strong> this background has contributed to the<br />

here developed approach: namely that aromatic peroxygenases (APOs) from the<br />

basidiomycetes A.aegerita (AaeAPO), C. radians (CraAPO) <strong>and</strong> M. rotula (MroAPO) can<br />

be used as an "artificial liver" to produce diverse HDMs.<br />

To this end, a representative selection <strong>of</strong> widely used <strong>pharmaceuticals</strong> <strong>and</strong> illicit <strong>drugs</strong> was<br />

tested with the above mentioned peroxygenases. In the first phase <strong>of</strong> this study, only<br />

AaeAPO <strong>and</strong> CraAPO were available, later also MroAPO was included in the in vitro<br />

experiments. The results have shown that all <strong>fungal</strong> peroxygenases (APOs) tested<br />

catalyzed the hydroxylation or dealkylation <strong>of</strong> diverse <strong>pharmaceuticals</strong> <strong>and</strong> <strong>drugs</strong> <strong>of</strong> abuse,<br />

in most cases with high regioselectivity. Moreover, MroAPO was also found to catalyze<br />

the conversion <strong>of</strong> some steroids that were not accessible to the <strong>other</strong> APOs. The findings<br />

<strong>of</strong> this study confirm previous work, which suggested that extracellular APOs share<br />

functional characteristics with intracellular P450s but utilize the peroxide-dependent shunt<br />

pathway with much higher efficiency (Guengerich 2001, H<strong>of</strong>richter <strong>and</strong> Ullrich 2010,<br />

Ortiz de Montellano <strong>and</strong> de Voss 2005, Shoji et al. 2007). These peroxygenations are<br />

thought to be initiated when the enzyme heme is oxidized <strong>by</strong> H 2 O 2 to give an iron species<br />

(oxo-ferryl iron, compound I) that carries one <strong>of</strong> the peroxide oxygens, which<br />

subsequently oxidizes a C-H bond in the substrate (Rittle <strong>and</strong> Green 2010).<br />

91


DISCUSSION<br />

Figure 37 Molecular structures <strong>of</strong> <strong>pharmaceuticals</strong> <strong>and</strong> <strong>drugs</strong> converted <strong>by</strong> AaeAPO. The substrates<br />

are divided into three different groups according to the relative effectiveness <strong>of</strong> their oxidation.<br />

Arrows indicate the points <strong>of</strong> enzymatic attack. SC – substrate conversion<br />

Figures 37-39 give an overview <strong>of</strong> pharmaceutically relevant peroxygenase substrates<br />

classified <strong>by</strong> the effectiveness <strong>of</strong> their conversion. What at first glance (Figure 37) attracts<br />

the attention is the better (60-100%) conversion <strong>by</strong> AaeAPO <strong>of</strong> relatively small aromatic<br />

substances or less complex molecules. However, more complex structures with exposed<br />

aromatic rings, with secondary or tertiary amino groups or methoxy groups were<br />

92


DISCUSSION<br />

nevertheless oxidized <strong>by</strong> the enzyme to extents ranging from 20 to 59%. Amphetamines<br />

were only hydroxylated at the aromatic ring <strong>and</strong> after long reaction times, <strong>and</strong> even then<br />

the conversion rate was relatively low (below 15%). CraAPO has a similar substrate<br />

spectrum as AaeAPO but the conversion rates were considerably lower in most cases<br />

(Figure 38). On the <strong>other</strong> h<strong>and</strong>, compared to AaeAPO, an overview <strong>of</strong> the MroAPO<br />

substrates suggests a preferance for more complex non-aromatic structures <strong>by</strong> the enzyme<br />

as found, for example, with steroids or codeine. The latter substrates were hardly or not at<br />

all oxidized <strong>by</strong> AaeAPO <strong>and</strong> CraAPO, which indicates considerable differences in the<br />

heme channels <strong>and</strong>/or the active sites <strong>of</strong> the enzymes.<br />

Furthermore, most <strong>of</strong> the observed oxidations matched those catalyzed <strong>by</strong> human liver<br />

P450s, e.g. propranolol (I) to 5-hydroxypropranolol (CYP2D6) (Masubuchi et al. 1994),<br />

tolbutamide (XV) to 4-hydroxytolbutamide (CYP2C9) (Miners <strong>and</strong> Birkett 1996),<br />

naproxen (XIX) to O-desmethylnaproxen (CYP1A2) (Miners et al. 1996), acetanilide (II)<br />

to acetaminophen (CYP1A2) (Liu et al. 1991, Miners et al. 1996) <strong>and</strong> sildenafil (XXIV) to<br />

N-desmethylsildenafil (CYP3A4) (Ku et al. 2008). Particularly interesting results will be<br />

discussed in detail in the next part <strong>of</strong> this chapter.<br />

Evidently, APOs have some advantages over P450s - including currently developed,<br />

laboratory-evolved P450s - where reaction yields are concerned (Kinne et al. 2009a). For<br />

example, the transformation <strong>of</strong> dicl<strong>of</strong>enac (IV) to 4-hydroxydicl<strong>of</strong>enac <strong>by</strong> mutants <strong>of</strong><br />

P450 cam (CYP101A1) resulted in total conversions between 15 <strong>and</strong> 44% <strong>and</strong> yields around<br />

10% (Weis et al. 2009). By contrast, the wild-type AaeAPO used here gave a total<br />

conversion <strong>of</strong> 78% <strong>and</strong> a yield <strong>of</strong> 68% for the same reaction. For the metabolism <strong>of</strong><br />

propranolol (II) in humans, CYP2D6 <strong>and</strong> CYP1A2 are responsible, forming three major<br />

products: 5-hydroxypropranolol, 4’-hydroxypropranolol <strong>and</strong> desisopropylpropranolol<br />

(Masubuchi et al. 1994). Positive results concerning the production <strong>of</strong> propranolol<br />

metabolites were reported for laboratory-evolved bacterial P450s that use the shunt<br />

pathway to hydroxylate the pharmaceutical. Thus, an engineered mutant D6H10 from<br />

P450 BM3 (Bacillus megaterium, expressed in E.coli) was optimized for the synthesis <strong>of</strong><br />

propranolol metabolites, especially the 5-hydroxylated product. The optimization resulted<br />

in a yield <strong>of</strong> 0.5% <strong>of</strong> 5- hydroxypropranolol (6.5 mg L -1 ) when 5 mM (1,295 mg L -1 )<br />

propranolol was converted (Otey et al. 2006). Analogous experiments with AaeAPO gave<br />

yields up to 13.6% (176 mg L -1 ), with regioselectivity higher than 90%. Interestingly,<br />

CraAPO produced the <strong>other</strong> human P450 metabolites <strong>of</strong> propranolol, too. The analysis<br />

93


DISCUSSION<br />

showed the formation <strong>of</strong> the following metabolites in the stated yields: 6% <strong>of</strong> 5-<br />

hydroxypropranolol, 2% <strong>of</strong> 4’-hydroxypropranolol <strong>and</strong> 3% <strong>of</strong> desisopropylpropranolol.<br />

The conversion <strong>of</strong> ibupr<strong>of</strong>en (XIV) is an<strong>other</strong> example for the formation <strong>of</strong> different<br />

metabolite ratios <strong>by</strong> different peroxygenases. In humans, more than 70% <strong>of</strong> ibupr<strong>of</strong>en is<br />

hydroxylated. The formation <strong>of</strong> metabolites is dependent on the P450 system <strong>and</strong><br />

comprises 2-hydroxy-, 3-hydroxy-, <strong>and</strong> 1-hydroxyibupr<strong>of</strong>en (Hamman et al. 1997, Holmes<br />

et al. 2007). AaeAPO converted ibupr<strong>of</strong>en to 2-hydroxyibupr<strong>of</strong>en <strong>and</strong> 1-hydroxyibupr<strong>of</strong>en<br />

with yields <strong>of</strong> 21% <strong>and</strong> 7%, respectively. The reaction with CraAPO showed high<br />

regioselectivity <strong>and</strong> gave only 2-hydroxyibupr<strong>of</strong>en with a yield <strong>of</strong> 72%.<br />

Not only hydroxylations but also N-dealkylations were differently catalyzed <strong>by</strong> different<br />

peroxygenases. For example, the oxidation <strong>of</strong> lidocaine (XXV) catalyzed <strong>by</strong> AaeAPO gave<br />

25% <strong>of</strong> monoethylglycinexylidide <strong>and</strong> 18% <strong>of</strong> glycinexylidide (ratio 1.4:1), whereas<br />

CraAPO formed 32% <strong>of</strong> monoethylglycinexylidide (with a regioselectivity <strong>of</strong> 70%) but<br />

only 5% <strong>of</strong> glycinexylidide (ratio 6.4:1). Both substances are major human metabolites<br />

when lidocaine is used as local anesthetic (Orl<strong>and</strong>o et al. 2004, Tam et al. 1990).<br />

The different specificities <strong>of</strong> <strong>fungal</strong> peroxygenases are an interesting aspect, suggesting<br />

that a set <strong>of</strong> them might be used as a "biocatalytic toolbox". Depending on what<br />

reaction/metabolite is needed, the appropriate enzyme could be chosen <strong>and</strong> applied under<br />

appropriate conditions. This interesting approach will be clarified <strong>by</strong> the following<br />

examples.<br />

The regioselective O-dealkylation <strong>of</strong> the anti-inflamatory drug naproxen (XIX) leads to the<br />

formation <strong>of</strong> 6-desmethylnaproxen as the only metabolite in the human liver (Segre 1975).<br />

In catalyzing this reaction, AaeAPO (yield <strong>of</strong> 57 %) was more effective than CraAPO (9%<br />

yield). On the <strong>other</strong> h<strong>and</strong>, CraAPO was the only peroxygenase to catalyze the unusual<br />

cleavage <strong>of</strong> an ester bond in the antiviral substance oseltamivir, which led to the formation<br />

<strong>of</strong> the human metabolite oseltamivir carboxylate (yield 71%; formed in humans <strong>by</strong> hepatic<br />

carboxylesterases; (Lindegardh et al. 2006, McClellan <strong>and</strong> Perry 2001), which is the<br />

bioactive form <strong>of</strong> the drug (McClellan <strong>and</strong> Perry 2001). Sildenafil (XXIV) is an inhibitor<br />

<strong>of</strong> the human cGMP-specific phosphodiesterase type 5 <strong>and</strong> widely used as an antiimpotence<br />

drug (Ballard et al. 1998, Boolell et al. 1996). The compound was N-<br />

dealkylated <strong>by</strong> all tested peroxygenases at the piperazine ring to form N-<br />

desmethylsildenafil, the major metabolite circulating <strong>and</strong> operating in vivo. (Walker 1999).<br />

The conversion rates were quite high in all cases (>80 %), but the yields <strong>of</strong> N-<br />

desmethylsildenafil differed noticeably with only 4% for CraAPO <strong>and</strong> almost 90% for<br />

94


DISCUSSION<br />

AaeAPO <strong>and</strong> MroAPO. Other pathways <strong>of</strong> the metabolism <strong>of</strong> sildenafil lead in humans to<br />

the formation <strong>of</strong> piperazine-N,N-deethylation <strong>and</strong> aliphatic hydroxylation products<br />

(Muirhead et al. 2002, Walker 1999).<br />

The benzodiazepine compounds diazepam (VIII), temazepam (IX) <strong>and</strong> oxazepam (X)<br />

used as psychoactive <strong>drugs</strong> in the treatment <strong>of</strong> mental deseases were efficiently<br />

hydroxylated (m/z +16) <strong>by</strong> AaeAPO.<br />

Figure 38 Molecular structures <strong>of</strong> <strong>pharmaceuticals</strong> <strong>and</strong> <strong>drugs</strong> converted <strong>by</strong> CraAPO. The<br />

substrates are divided into three different groups according to the relative effectiveness <strong>of</strong> their<br />

oxidation. Arrows indicate the points <strong>of</strong> enzymatic attack. SC – substrate conversion<br />

95


DISCUSSION<br />

The reaction <strong>of</strong> diazepam with human P450s gives also a hydroxylated derivative as the<br />

major metabolite (temazepam, hydroxylated at the heterocyclic 7-ring) <strong>and</strong> N-<br />

demethylated diazepam (N-desmethyldiazepam) as a second metabolite (Andersson et al.<br />

1994). However, since the hydroxylated product <strong>of</strong> the AaeAPO reaction did not have the<br />

same retention time as temazepam, <strong>and</strong> taking into account the preference <strong>of</strong> AaeAPO for<br />

aromatic substrates (Ar<strong>and</strong>a et al. 2009a, Ar<strong>and</strong>a et al. 2009b), the hydroxylation <strong>of</strong> one <strong>of</strong><br />

the carbons <strong>of</strong> the phenyl moiety would be a plausible explanation (most probably in parapostion).<br />

Temazepam <strong>and</strong> oxazepam were seemingly hydroxylated <strong>by</strong> AaeAPO in a<br />

similar way, <strong>and</strong> in the case <strong>of</strong> the latter compound, also a double hydroxylated product<br />

was detected (m/z +32). By contrast, P450s preferably N-demethylate temazepam to form<br />

oxazepam (Schwarz 1979), which in turn is almost exclusively conjugated with glucuronic<br />

acid to improve excretion (Kellermann <strong>and</strong> Luyten-Kellermann 1979).<br />

The antitussive morphine-derivative codeine (XXVIII), which was not a substrate <strong>of</strong><br />

AaeAPO or CraAPO, was rapidly converted <strong>by</strong> MroAPO into the N-dealkylated product<br />

norcodeine <strong>and</strong> traces <strong>of</strong> the O-dealkylation product morphine (XXIX). Unexpectedly, no<br />

product was found when morphine was treated with MroAPO under st<strong>and</strong>ard conditions;<br />

only in the absence <strong>of</strong> ascorbic acid, a reaction was observed leading to the formation <strong>of</strong> a<br />

coupling product (dimer, m/z 569) that was later also observed for AaeAPO <strong>and</strong> CraAPO.<br />

This fact indicates that peroxygenases attack the phenolic group <strong>of</strong> morphine via oneelectron<br />

oxidation resulting in the formation <strong>of</strong> a phenoxyl radical that in turn can dimerize<br />

(Figure 22, Figure 50). Since the morphine molecule <strong>and</strong> its coupling product are relatively<br />

bulky, the oxidation may occur at the enzyme surface rather than in the heme channel.<br />

P450s (e.g. human CYP2D6) preferably O-demethylate codeine to form morphine in Phase<br />

I metabolism, <strong>and</strong> this product is in turn conjugated with glucuronic acids <strong>and</strong> excreted<br />

(Kilpatrick <strong>and</strong> Smith 2005, Williams et al. 2002). Tests with human plasma samples<br />

showed that codeine is primarily conjugated at the 6-position. The P450-catalyzed<br />

demethylation clearance rates <strong>of</strong> codeine <strong>and</strong> norcodeine were found to be similar in<br />

magnitude but about four times lower compared to 6-glucuronidation. Since the partial<br />

clearance to norcodeine was normal in poor human metabolisers, it would be reasonable to<br />

infer that the two demethylation processes (N-demethylation to norcodeine, O-<br />

demethylation to morphine) are mediated <strong>by</strong> different cytochrome P-450 isoenzymes<br />

(Chen et al. 1991). In contrast to morphine, the psychoactive, illicit drug<br />

tetrahydrocannabinol (THC, XVIII) was efficiently converted <strong>by</strong> AaeAPO <strong>and</strong> CraAPO<br />

(conversion rate almost 100%) resulting in the formation <strong>of</strong> two hydroxylated products, the<br />

96


DISCUSSION<br />

major <strong>of</strong> which is probably 11-hydroxy-THC; proposed structures are shown in the results<br />

section (Figure 19).<br />

Figure 39 Molecular structures <strong>of</strong> <strong>pharmaceuticals</strong> <strong>and</strong> <strong>drugs</strong> converted <strong>by</strong> MroAPO. The<br />

substrates are divided into three different groups according to the relative effectiveness <strong>of</strong> their<br />

oxidation. Arrows indicate the points <strong>of</strong> enzymatic attack. SC – substrate conversion.<br />

In <strong>other</strong> experiments where side chains were attacked the best results were noted for<br />

MroAPO but in the case <strong>of</strong> THC, no positive result was observed. In the human liver, THC<br />

is metabolized mainly to 11-hydroxy-THC, which is further oxidized to the corresponding<br />

carboxylic acid (11-nor-9-carboxy-THC); both reactions are thought to be catalyzed <strong>by</strong><br />

P450s (CYP2C9, CYP2C19, CYP3A4) (Bl<strong>and</strong> et al. 2005, Lemberger 1973, Watanabe et<br />

al. 2007).<br />

97


DISCUSSION<br />

None <strong>of</strong> the steroids tested were attacked <strong>by</strong> CraAPO <strong>and</strong> AaeAPO; however, MroAPO<br />

again was able to convert these bulky molecules. According to the effectiveness <strong>of</strong> the<br />

conversion, two groups <strong>of</strong> steroid substrates can be distinguished. Substrates <strong>of</strong> the first<br />

group were converted <strong>by</strong> more than 95% (e.g. Reichstein's substance S, cortisone,<br />

prednisone, medrol), while those <strong>of</strong> the second group were oxidized with lower rates (2-<br />

70% conversion; e.g. progesterone, testosterone, n<strong>and</strong>rolone, pregnenolone). All steroids<br />

<strong>of</strong> the first group bear a 1-oxo-2-hydroxyethyl side chain (-CO-CH 2 -OH), whereas those <strong>of</strong><br />

the second group have a 1-oxoethyl (-CO-CH 3 ) or shorter (e.g.<br />

-CH 3 ) side chain, or no side chain at all (e.g. -OH). This finding indicates that the primary<br />

alcohol functionality <strong>of</strong> the oxohydroxyethyl group somehow facilitates the conversion <strong>of</strong><br />

steroids <strong>by</strong> MroAPO (<strong>and</strong> maybe it is also the target <strong>of</strong> oxidation). Unfortunately, the mass<br />

spectral data <strong>of</strong> the products are difficult to interpret, so that incorporation <strong>of</strong> oxygen (m/z<br />

+16) could be deduced in only a few cases, an (e.g. for Reichstein's substance S,<br />

pregnenolone, progesterone; see also Table 18 in the results section). Nevertheless the<br />

mass data indicate an<strong>other</strong> interesting reaction that led to the formation <strong>of</strong> a progesteronelike<br />

product (m/z 315) from pregnenolone (m/z 317). The reaction could proceed via a<br />

second hydroxylation <strong>of</strong> the C3-position yielding a geminal alcohol that eliminates water<br />

to form a keto group. However, it is unclear whether <strong>and</strong> if so how the double bound is<br />

transferred from C5 to C4. In animals <strong>and</strong> humans these reactions are catalyzed <strong>by</strong> the 3-βhydroxysteroid<br />

dehydrogenase/Δ-4-5 isomerase (EC 1.1.1.145), the only enzyme in the<br />

adrenal pathway <strong>of</strong> steroid synthesis that is not a member <strong>of</strong> the P450 superfamily<br />

(Thomas et al. 2004). Various <strong>other</strong> human P450s are involved in steroid transformation<br />

<strong>and</strong> the synthesis <strong>of</strong> hormones; among <strong>other</strong>s, they hydroxylate different rings (e.g.<br />

CYP11B1, CYP17), the side chain (e.g. CYP21) <strong>and</strong> the C18-methyl group (CYP11B2),<br />

<strong>and</strong> as well they are involved in the aromatization <strong>of</strong> the A ring, there<strong>by</strong> converting<br />

<strong>and</strong>rogens to estrogens (CYP19) (Hall 1986).<br />

The results <strong>of</strong> steroid conversion clearly display some limitations <strong>and</strong> drawbacks <strong>of</strong> APO<br />

catalysis. One is that the enzymes have relatively narrow catalytic clefts (albeit the hemechannel<br />

is considerably wider than those <strong>of</strong> <strong>other</strong> heme peroxidases), which prevent access<br />

<strong>of</strong> markedly bulky substrates to the active site (Piontek et al. 2010). The negative results<br />

we obtained when treating felbamate (LIX), 17α-ethinylestradiol (XLII), testosterone<br />

(XLI), Reichstein's substance S (XLIII) <strong>and</strong> <strong>other</strong> steroid structures with AaeAPO or<br />

CraAPO may reflect this limitation. Furthermore, all three tested peroxygenases were<br />

98


DISCUSSION<br />

unable to mimic some characteristic reactions <strong>of</strong> human P450s such as felodipine (LIX)<br />

<strong>and</strong> coumarin (LVII) oxidation. Why the small <strong>and</strong> hydrophobic coumarin molecule was<br />

not attacked remains an enigma. Although MroAPO had no problems with the conversion<br />

<strong>of</strong> steroids, it was not able to catalyze the hydroxylation <strong>of</strong> benzodiazepines, which on the<br />

<strong>other</strong> h<strong>and</strong> were good substrates <strong>of</strong> AaeAPO <strong>and</strong> CraAPO.<br />

In conclusion, <strong>fungal</strong> APOs may be better suited than most P450s for synthetic<br />

applications: (i) they utilize a relatively cheap co-substrate, H 2 O 2 ; (ii) they do not require<br />

costly co-reactants such as pyridine nucleotides, flavin reductases, or ferredoxins; (iii) they<br />

are secreted enzymes <strong>and</strong> thus can be cost-effectively produced (no cell disruption<br />

required); <strong>and</strong> (iv) they are stable <strong>and</strong> soluble due to their high degree <strong>of</strong> glycosylation<br />

(Pecyna et al. 2009, Ullrich et al. 2009).<br />

4.2 Mechanistic studies<br />

4.2.1 Source <strong>of</strong> oxygen <strong>and</strong> stoichiometry.<br />

The results <strong>of</strong> this work <strong>and</strong> the previous studies with APOs showed that an oxygen atom<br />

from hydrogen peroxide is incorporated into the phenolic or alcoholic functionalities<br />

produced after conversion <strong>of</strong> alkyl aromatics (Ar<strong>and</strong>a et al. 2009a, Kinne et al. 2009a,<br />

Kinne et al. 2008, Kluge et al. 2009), <strong>and</strong> also into the carbonyls produced after oxidation<br />

<strong>of</strong> alcohols or oxidative cleavage <strong>of</strong> ethers (Kinne et al. 2009b). According to the above<br />

results, oxygen incorporation from H 2 O 2 should be quantitative when the substrate is<br />

oxidized. The data on APO-catalyzed oxidation <strong>of</strong> <strong>pharmaceuticals</strong> agree with this picture,<br />

since almost 100% <strong>of</strong> oxygen present in newly generated phenolic or benzylichydroxylation<br />

products - e.g. in the phenolic group <strong>of</strong> acetaminophen or in the alcohol<br />

moiety <strong>of</strong> 4-hydroxytolbutamide - was 18 O-labeled when the experiment was conducted<br />

with H 18 2 O 2 . This result correlated well with the reaction mechanism reported for<br />

engineered P450 BM3 or P450 cam , which can use the so-called “peroxide shunt” pathway,<br />

in the course <strong>of</strong> which also one oxygen atom from hydrogen peroxide is incorporated into<br />

the substrate to yield an oxygenated product without the need <strong>of</strong> a reductase domain,<br />

dioxygen or NADPH; (Joo et al. 1999, Otey et al. 2006). Moreover, the stoichiometrical<br />

experiment with sildenafil showed that one equivalent <strong>of</strong> N-desmethylsildenafil was<br />

formed per equivalent <strong>of</strong> H 2 O 2 supplied. This agrees with the two-electron oxidation<br />

expected from a peroxygenative mechanism (Section 1.2.7; (Gorsky et al. 1984)). The<br />

99


DISCUSSION<br />

stoichiometrical experiment with benzodioxoles seems to follow an<strong>other</strong> mechanism <strong>and</strong><br />

that is described in one <strong>of</strong> the following subchapters (4.2.7).<br />

4.2.2 Deuterium isotope effect<br />

Historically, the determination <strong>of</strong> deuterium isotope effects has been a powerful tool to<br />

help unravel the intricacies <strong>of</strong> carbon-hydrogen bond cleavage <strong>and</strong> define the mechanism<br />

<strong>of</strong> specific chemical reactions. The intrinsic primary isotope effect, k H /k D , is the magnitude<br />

<strong>of</strong> the isotope effect for a reaction on the rate constant for the bond-breaking step (C–H<br />

versus C–D) <strong>of</strong> the reaction <strong>and</strong> is related to the symmetry <strong>of</strong> the transition state for that<br />

step (Nelson <strong>and</strong> Trager 2003). There are two designs to follow deuterium isotope effects.<br />

Intermolecular design depends on complex determinations <strong>of</strong> rate constants, whereas<br />

intramolecular design is kinetically independent <strong>of</strong> all steps besides the C-H bond-breaking<br />

step, <strong>and</strong> is there<strong>by</strong> simpler. In an isotope effect experiment <strong>of</strong> intramolecular design, a<br />

substrate that is susceptible to enzymatic attack at either <strong>of</strong> two symmetrically equivalent<br />

sites is chosen (Hjelmel<strong>and</strong> et al. 1977, Nelson <strong>and</strong> Trager 2003). One site contains<br />

deuterium <strong>and</strong> the <strong>other</strong> retains its normal complement <strong>of</strong> hydrogen. The enzyme then has<br />

the choice <strong>of</strong> attacking a deuterated or an equivalent protio site within the same molecule.<br />

The observed isotope effect (k H /k D ) obs is measured as the ratio <strong>of</strong> product resulting from<br />

attack at the protio site versus product resulting from the attack at the deutero site (Kinne<br />

2010, Nelson <strong>and</strong> Trager 2003). The intramolecular deuterium isotope effect that was<br />

observed for phenacetin-d1 oxidation <strong>by</strong> AaeAPO, is consistent with a P450-like<br />

mechanism. The determined value <strong>of</strong> (k H /k D ) obs around 3 is close to the values <strong>of</strong> (k H /k D ) obs<br />

near 2 that have been observed for the P450-catalyzed O-dealkylation <strong>of</strong> this substrate<br />

(Garl<strong>and</strong> et al. 1976). A value <strong>of</strong> 2 is considerably smaller than the intrinsic isotope effect<br />

near 10 expected for a hydrogen abstraction mechanism, <strong>and</strong> may indicate that phenacetin<br />

oxidation <strong>by</strong> APOs may proceed instead via electron transfers, as proposed earlier for the<br />

P450-catalyzed reaction (Yun et al. 2000). By contrast, the O-dealkylation <strong>of</strong> 1,4-<br />

dimethoxybenzene-d 3 <strong>by</strong> AaeAPO probably does proceed via hydrogen abstraction,<br />

because it was found to exhibit a much higher (k H /k D ) obs near 12 (Groves <strong>and</strong> McClusky<br />

1978, Kinne et al. 2009b, Yun et al. 2000).<br />

4.2.3 Kinetics<br />

The apparent kinetic data determined in this study for the action <strong>of</strong> AaeAPO on a variety<br />

<strong>of</strong> <strong>pharmaceuticals</strong> suggest that peroxygenases may be a useful alternative to P450s for the<br />

100


DISCUSSION<br />

regioselective preparation <strong>of</strong> human drug metabolites (HDMs). Although some P450s are<br />

known to bind pharmaceutical substrates more strongly than APOs, exhibiting K m values<br />

between 1 <strong>and</strong> 70 µM, they generally exhibit relatively low k cat values in the vicinity <strong>of</strong> 0.2<br />

s -1 or less (Upthagrove <strong>and</strong> Nelson 2001, Yun et al. 2000). As a result, the catalytic<br />

efficiencies [k cat /K m ] <strong>of</strong> AaeAPO for pharmaceutical oxidations - despite relatively high K m<br />

values - lie in the same range as those <strong>of</strong> the P450s, as exemplified <strong>by</strong> our values for<br />

propranolol hydroxylation (1.32 × 10 5 M -1 s -1 ), metoprolol O-dealkylation (4.1 × 10 4 M -1 s -<br />

1 ) <strong>and</strong> phenacetin O-dealkylation (3.3 × 10 4 M -1 s -1 ). Moreover, the k cat /K m value we<br />

obtained for acetanilide (1.5 × 10 6 M -1 s -1 ) is much higher than the corresponding value<br />

found for the P450-catalyzed reaction (Liu et al. 1991).<br />

4.2.3.1 Bisubstrate kinetics<br />

In some cases, the kinetic investigation <strong>of</strong> APO-catalyzed reactions has encountered some<br />

experimental problems. For example, a previous study showed that, under saturating H 2 O 2<br />

concentrations, AaeAPO exhibits an interfering catalase activity, which can lead to an<br />

underestimation <strong>of</strong> kinetic efficiency (Ullrich et al. 2008). A better option for the<br />

evaluation <strong>of</strong> kinetic parameters would be bisubstrate systems using an experimental<br />

design that varies the concentration <strong>of</strong> both substrates (Segel 1994). Initial velocity<br />

patterns are usually obtained <strong>by</strong> making reciprocal plots for one substrate (variable<br />

substrate) at different fixed concentrations <strong>of</strong> one <strong>of</strong> the <strong>other</strong>s (changing fixed substrate),<br />

while keeping all <strong>other</strong> substrates, if there are any, at constant concentration. If the<br />

mechanism does not include alternate reaction sequences <strong>and</strong> no substrates react with more<br />

than one enzyme form, the intercepts always have to be a linear function <strong>of</strong> the reciprocal<br />

concentration <strong>of</strong> the changing fixed substrate. As a result, there are two possible initial<br />

velocity patterns: parallel lines when no reversible connection exists, or lines intersecting<br />

to the left <strong>of</strong> the vertical axis when such a connection does exist (Clel<strong>and</strong> 1963). Data from<br />

bisubstrate kinetics under steady state conditions, which were spectrophotometrically<br />

recorded for the AaeAPO-catalyzed reactions at varying concentrations <strong>of</strong> 4-<br />

nitrobenzodioxole <strong>and</strong> H 2 O 2, gave in fact parallel lines for double reciprocal plots <strong>of</strong> the<br />

data. This result is consistent with an enzymatic ping-pong mechanism (Bisswanger 2000,<br />

Clel<strong>and</strong> 1963, Segel 1994). The mechanism can be explained as follows. The substrate (A)<br />

binds to the free enzyme (E) to form the enzyme-substrate complex (EA). Substrate (A)<br />

converts the enzyme to the modified enzyme-product complex (FP). Only after the<br />

substrate (A) is released in the form <strong>of</strong> (P) can the second substrate (B) bind to the<br />

101


DISCUSSION<br />

modified enzyme (F), yielding the enzyme-substrate complex (FB), <strong>and</strong> then react with the<br />

modified enzyme, finally regenerating the free enzyme (E) with release <strong>of</strong> the product (Q)<br />

(Bisswanger 2000, Kinne 2010, Segel 1994).<br />

Figure 40 Reaction sequence <strong>of</strong> a "ping-pong bi bi" mechanism. The enzyme constantly bounces<br />

back <strong>and</strong> forth between the two states <strong>of</strong> the enzyme (E) <strong>and</strong> (F); substrate (A), second substrate<br />

(co-substrate) (B), product (Q), enzyme-substrate complex (EA), complex <strong>of</strong> the modified enzyme<br />

<strong>and</strong> the second substrate (FB), modified enzyme-product complex (FP), modified enzyme (F)<br />

The kinetic parameters determined in this way for the demethylenation <strong>of</strong> the model 4-<br />

nitrobenzodioxole <strong>by</strong> AaeAPO were in the same range as the data obtained for selected<br />

<strong>pharmaceuticals</strong>. Thus, the K m for 4-nitrobenzodioxole <strong>of</strong> 1,190 µM is between the value<br />

calculated for the N-demethylation <strong>of</strong> phenacetin (XXI; K m =999) <strong>and</strong> the hydroxylation <strong>of</strong><br />

acetanilide (II; K m = 1,310 µM). The K m for the demethylenation <strong>of</strong> methylenedioxymethamphetamine<br />

(XXXVII) catalyzed <strong>by</strong> P450s is <strong>by</strong> two orders <strong>of</strong> magnitude lower<br />

(K m =1.7-3 µM) (Tucker et al. 1994). The catalytic efficiency for nitrobenzodioxole O-<br />

demethylenation (k cat /K m = 3.81×10 5 M -1 s -1 ), lies in the same range as that for<br />

naphthalene oxidation (k cat /K m = 5,2 × 10 5 M -1 s -1 ; see Chapter 4.2.3 <strong>and</strong> 3.9.1).<br />

4.2.4 Assays for human cytochrome P450 activities<br />

Twelve validated assays for human hepatic cytochrome P450s have been established in the<br />

last decades. Using this set <strong>of</strong> assays <strong>and</strong> substrates, six subfamilies, including nine<br />

individual P450s can be characterized (Anzenbacher <strong>and</strong> Anzenbacherová 2001, Walsky<br />

<strong>and</strong> Obach 2004, Wrighton <strong>and</strong> Stevens 1992). All twelve assays were applied here for the<br />

identification <strong>of</strong> specific P450-like activities <strong>of</strong> three <strong>fungal</strong> peroxygenases (AaeAPO,<br />

CraAPO <strong>and</strong> MroAPO; see Table 22, Chapter 3.12).<br />

In the liver, CYP1A2 metabolizes about 15% <strong>of</strong> all incoming <strong>pharmaceuticals</strong> (Gibson <strong>and</strong><br />

Skett 2008). The activity <strong>of</strong> this enzyme is measured with the phenacetin (XXI) O-<br />

deethylase assay; only MroAPO was unable to perform this reaction. Two additional<br />

substrates <strong>of</strong> CYP1A2 <strong>and</strong> AaeAPO/CraAPO are naproxen (XIX) <strong>and</strong> 3,4-<br />

aminoantipyrine (XXVI) (Anzenbacher <strong>and</strong> Anzenbacherová 2001). CYP2A6 contributes<br />

102


DISCUSSION<br />

about 5% <strong>of</strong> cytochrome P450-dependent drug metabolism in the liver. One <strong>of</strong> its<br />

characteristic catalytic properties, a coumarin 7-hydroxylase activity, is seemingly lacking<br />

in APOs. Other substrates <strong>of</strong> CYP2A6 are biogenic amines including tobacco-specific<br />

nitrosamines with pro-carcinogenic potential (Guengerich <strong>and</strong> Montellano 2005).<br />

However, CYP2A6 is generally considered less important in drug metabolism than <strong>other</strong><br />

P450s (Walsky <strong>and</strong> Obach 2004). CYP2B6 is responsible for only about 1% <strong>of</strong> hepatic<br />

metabolism <strong>of</strong> <strong>pharmaceuticals</strong>. This enzyme is characterized <strong>by</strong> its bupropione<br />

hydroxylase activity. Although AaeAPO <strong>and</strong> CraAPO converted bupropione to<br />

hydroxylated products, the absence <strong>of</strong> an analytical st<strong>and</strong>ard did not allow a clear<br />

conclusion as to whether the same carbon is attacked. An<strong>other</strong> marker reaction <strong>of</strong> CYP2B6<br />

is the N-demethylation <strong>of</strong> S-mephenytoin, which was performed <strong>by</strong> the APOs only to a<br />

low extent. CYP2C8 converts less than 5% <strong>of</strong> incoming <strong>pharmaceuticals</strong> in the liver <strong>and</strong><br />

has a generally low catalytic activity towards the known substrates <strong>of</strong> the CYP2C<br />

subfamily (Guengerich <strong>and</strong> Montellano 2005); its characteristic amodiaquine N-<br />

deethylase activity was detected for all APOs as well. CYP2C9 is involved in the<br />

metabolism <strong>of</strong> about 20% <strong>of</strong> <strong>pharmaceuticals</strong> <strong>and</strong> <strong>drugs</strong>, <strong>and</strong> is one <strong>of</strong> the major enzymatic<br />

metabolizers in the liver (Guengerich <strong>and</strong> Montellano 2005). It is characterized <strong>by</strong><br />

hydroxylating activities towards dicl<strong>of</strong>enac <strong>and</strong> tolbutamide, both excellent substrates <strong>of</strong><br />

AaeAPO <strong>and</strong> also convertable <strong>by</strong> the <strong>other</strong> APOs (but with lower yields). Additional<br />

substrates are carbamazepine (III) sildenafil (XXIV) (Warrington et al. 2000) <strong>and</strong><br />

ibupr<strong>of</strong>en (XIV). S-mephenytoin 4’-hydroxylation is the activity characteristic for<br />

CYP2C19, which participates in less than 5% <strong>of</strong> hepatic drug metabolism. APOs<br />

seemingly share this activity with CYP2C19, <strong>and</strong> also the ability to efficiently oxidize<br />

diazepam (VIII), omeprazole (XVII) <strong>and</strong> propranolol (I). The CYP2E subfamily is present<br />

in all mammals in the form <strong>of</strong> the characteristic representative CYP2E1. This enzyme is<br />

well-known for its involvement in the metabolism <strong>of</strong> ethanol (Quertemont 2004, Zimatkin<br />

<strong>and</strong> Deitrich 1997). Concerning pharmaceutical <strong>drugs</strong>, it was shown to convert<br />

acetaminophen, phenacetin (XXI) <strong>and</strong> chlorzoxazone (LIV), the latter <strong>of</strong> which is the<br />

specific assay substrate for CYP2E1 (Anzenbacher <strong>and</strong> Anzenbacherová 2001).<br />

Chlorzoxazone (LIV) was found to be a good substrate <strong>of</strong> AaeAPO <strong>and</strong> CraAPO but, in<br />

the absence <strong>of</strong> an appropriate analytical st<strong>and</strong>ard, it was not possible to confirm that the<br />

same product was formed. The CYP3A subfamily with four human members contributes<br />

to about 50% <strong>of</strong> the hepatic metabolism <strong>of</strong> <strong>drugs</strong> <strong>and</strong> <strong>pharmaceuticals</strong> (Anzenbacher <strong>and</strong><br />

Anzenbacherová 2001, Guengerich <strong>and</strong> Montellano 2005). The substrate spectrum <strong>of</strong> these<br />

103


DISCUSSION<br />

enzymes is very broad <strong>and</strong> includes, inter alia, various statins, steroids, <strong>and</strong> codeine<br />

(XXVIII), but also APO substrates such as sildenafil (XXIV), lidocaine (XXV),<br />

acetaminophen, carbamazepine (III), dextromethorphan (XX, also converted <strong>by</strong><br />

CYP2D6), diazepam (VIII), omeprazole (XXII), <strong>and</strong> tamoxifen (XXVII). The typical<br />

CYP3A substrates felodipine <strong>and</strong> midazolam (both belonging to the twelve key assay<br />

substrates) were not at all or only slowly converted <strong>by</strong> the APOs. An<strong>other</strong> CYP3A<br />

substrate, testosterone, was only oxidized <strong>by</strong> MroAPO.<br />

The results <strong>of</strong> this work clearly show that APOs share catalytic activities with all relevant<br />

human P450s. Since the substrate spectra <strong>of</strong> P450s overlap with each <strong>other</strong>, it is not<br />

possible to select a particular P450 enzyme that shows the most similarities with APOs at<br />

the reaction level. Some limitations <strong>of</strong> APO catalysis <strong>and</strong> differences with human P450s<br />

were also acertained (coumarin <strong>and</strong> felodipine were converted <strong>by</strong> none <strong>of</strong> the APOs,<br />

testosterone only to some extent <strong>by</strong> MroAPO), <strong>and</strong> may be attributable to issues <strong>of</strong><br />

molecule size or charge, <strong>and</strong>/or special features <strong>of</strong> the 3-dimensional structure <strong>of</strong> the<br />

substances (see 4.5 for more details). Table 38 summarizes the results <strong>of</strong> this study.<br />

Table 38 Summary <strong>of</strong> the measured activities <strong>by</strong> P450s <strong>and</strong> APOs; +(red)- assay metabolite was<br />

found; (+)(orange)- metabolite with the same ion mass was found, but analitycal st<strong>and</strong>ard was<br />

lacking; *- N-dealkylation; **- hydroxylation<br />

4.2.5 Peroxides <strong>and</strong> peroxide generation systems<br />

In the literature, there are many ways described to add or generate peroxides to initiate<br />

peroxidation/peroxygenation reactions. As a part <strong>of</strong> this work, three different hydrogen<br />

peroxide generating systems <strong>and</strong> three peroxides were tested in the AaeAPO-catalyzed<br />

conversion <strong>of</strong> dicl<strong>of</strong>enac to 4'-hydroxydicl<strong>of</strong>enac.<br />

One approach that goes back to studies in the 1950s uses the the oxidation <strong>of</strong> ascorbic acid<br />

<strong>by</strong> dioxygen in the presence <strong>of</strong> Cu(II) perchlorate or L-cysteine, which generates<br />

104


DISCUSSION<br />

dehydroascorbic acid <strong>and</strong> hydrogen peroxide (Davies 1984, Khan <strong>and</strong> Martell 1967,<br />

Udenfriend et al. 1954). Though the details <strong>of</strong> H 2 O 2 formation are not fully clear, two<br />

possible mechanisms may be considered for metal-catalyzed peroxide formation via<br />

ascor<strong>by</strong>l radical intermediates. In the first proposed mechanism a metal-ascorbate complex<br />

is suggested in a pre-equilibrium step. This is followed <strong>by</strong> a rate-determining electron<br />

transfer within the complex from the ascorbate anion to the metal ion. The reduced<br />

complex dissociates in a fast step to the lower valence metal ion <strong>and</strong> a semiquinone.<br />

Afterwards follows reoxidation <strong>of</strong> the reduced metal ion <strong>by</strong> O 2 to produce superoxide,<br />

which then generates H 2 O 2 via dismutation.<br />

•<br />

2<br />

fast<br />

2HO<br />

⎯⎯→H<br />

O + O<br />

2<br />

2<br />

2<br />

In the second suggested mechanism, at first a metal-ascorbate complex is formed. This is<br />

followed <strong>by</strong> an attack <strong>of</strong> dioxygen on the metal-ascorbate complex with the formation <strong>of</strong><br />

an oxygen complex in a second pre-equilibrium step (Khan <strong>and</strong> Martell 1967).<br />

AscH<br />

M<br />

−<br />

( n−1)<br />

+<br />

+ O ⎯⎯→ Asc + HO<br />

+ fast<br />

+ HO H ⎯⎯→M<br />

fast<br />

2<br />

•<br />

2<br />

+<br />

•<br />

2<br />

n+<br />

+ H<br />

In the case <strong>of</strong> the cysteine system, it is thought that the initial oxidation <strong>of</strong> the thiolate<br />

group <strong>of</strong> the amino acid into a thiyl radical initiates peroxide generation (Law et al.<br />

1983) 11 . An<strong>other</strong> peroxide-generating system, which has widely been used in enzymatic<br />

conversions for decades <strong>and</strong> was also successfully tested here, is that <strong>of</strong> glucose/glucose<br />

oxidase (H<strong>of</strong>richter et al. 1998, Paice et al. 1993, Ullrich et al. 2004). tert-<br />

Butylhydroperoxide <strong>and</strong> meta-chloroperoxybenzoic acid (mCPBA) were tested as organic<br />

alternatives replacing hydrogen peroxide in this work. The former peroxide was shown to<br />

be the best oxidant in sulfur/sulfide <strong>and</strong> olefin oxidation <strong>by</strong> chloroperoxidase <strong>and</strong><br />

horseradish peroxidase (Colonna et al. 1990, Hu <strong>and</strong> Hager 1999). Also mCPBA was used<br />

as an oxidant in many previous studies, particularly when the oxidations with H 2 O 2 were<br />

too fast (Kellner et al. 2002, MacKeith et al. 1994).<br />

Recent studies with hydrogen peroxide have already shown that the dosage <strong>and</strong> the way <strong>of</strong><br />

adding it can considerably influence the efficiency <strong>of</strong> a APO-catalyzed reaction (Ar<strong>and</strong>a et<br />

al. 2009b, Poraj-Kobielska et al. 2011b). The results obtained here demonstrated that<br />

hydrogen peroxide, mCPBA, <strong>and</strong> the system glucose/glucose oxidase are the best oxidants<br />

2<br />

O<br />

2<br />

11 Instead glutathione described in this paper, cysteine was used (Scheibner <strong>and</strong> H<strong>of</strong>richter 1998).<br />

105


DISCUSSION<br />

in terms <strong>of</strong> the dicl<strong>of</strong>enac conversion rates (32-45%). A stepwise dosage <strong>of</strong> the oxidant<br />

(with a pipette or a syringe pump) further increased product formation (<strong>by</strong> 3-7%) but was<br />

not nessesarily required. The worst results were observed for the L-cysteine/ascorbate<br />

system followed <strong>by</strong> the Cu(II) chlorate/ascorbate system <strong>and</strong> the usage <strong>of</strong> tertbutylhydroperoxide<br />

(conversion rates for dicl<strong>of</strong>enac 6-13%). This experiment showed that<br />

different peroxides <strong>and</strong> sources <strong>of</strong> them can be used in APO-catalyzed reactions; the<br />

effectiveness depends on the particular peroxide, its dosage <strong>and</strong> the reaction time.<br />

4.2.6 Difference binding spectra<br />

One <strong>of</strong> the characteristics <strong>of</strong> P450s is the influence <strong>of</strong> substrate binding on the appearance<br />

<strong>of</strong> the overall spectrum, especially with respect to the positions <strong>and</strong> intensities <strong>of</strong> the major<br />

absorption b<strong>and</strong>s (Lewis 2001, Schenkman et al. 1981). AaeAPO shows difference binding<br />

spectra that are similar to those <strong>of</strong> P450s. The <strong>pharmaceuticals</strong> tested with AaeAPO in this<br />

respect were: propranolol (II), dicl<strong>of</strong>enac (IV) <strong>and</strong> dextromethorphan (XX), as well as<br />

phenol (C 6 H 5 -OH) as a reference (Kinne 2010, Lewis 2001). They all gave a binding type I<br />

spectrum that is also characteristic for the binding exhibited <strong>by</strong> phenol (Hayhurst et al.<br />

2001, Schenkman et al. 1981, Topham 1970, Wester et al. 2003). These results suggest<br />

that the electronic environment <strong>of</strong> APOs, especially <strong>of</strong> AaeAPO, is similar to that <strong>of</strong><br />

P450s, although the amino acid residues in the heme channel <strong>and</strong> near the heme may be<br />

totally different.<br />

4.2.7 Phenoxyl radical scavenger<br />

The high general ("classic") peroxidase activity <strong>of</strong> APOs can cause problems when the<br />

desired products are phenols (e.g. after aromatic hydroxylation or O-demethylation <strong>of</strong><br />

methoxyaromatics), <strong>and</strong> necessitates the inclusion <strong>of</strong> a radical scavenger such as ascorbate<br />

in the reactions.The peroxidase activity <strong>of</strong> APOs causes the formation <strong>of</strong> phenoxyl radicals<br />

from phenolic compounds, which leads to undesired coupling reactions. Ascorbic acid is<br />

present in an aqueous environment in a protonated form (at pH 7) as a monoanionic<br />

species. The latter (probably as metal ion-ascorbate anion complex, see 1.2.5) can react<br />

with a phenoxyl radical to give the ascor<strong>by</strong>l radical that in turn may disproportionate to<br />

dehydroascorbic acid <strong>and</strong> ascorbic acid. In <strong>other</strong> words, the phenoxyl radical<br />

(R-C 6 H 5 -O • ) abstracts one electron from ascorbate followed <strong>by</strong> hydrogen/proton rebound<br />

to reform the phenolic compound (R-C 6 H 5 -OH). Figure 41 shows the proposed phenoxyl<br />

radical cycle in the presence <strong>of</strong> ascorbic acid with propranolol <strong>and</strong> 5-hydroxypropranolol<br />

106


DISCUSSION<br />

as APO substrate <strong>and</strong> phenolic product, respectively (H<strong>of</strong>richter et al. 2010, Kinne et al.<br />

2009a, Poraj-Kobielska et al. 2011b).<br />

Figure 41 Ascorbic acid as phenoxyl radical scavenger. The radical scavenging cycle is illustrated<br />

with propranolol as APO substrate. AscH 2 -ascorbic acid, AscH - -monoprotonated (monoionic)<br />

species <strong>of</strong> ascorbic acid, Asc 2- -diionic species, Asc •- -ascor<strong>by</strong>l radical, Asc- dehydroascorbic acid.<br />

4.2.8 Stoichiometry <strong>and</strong> proposed mechanism <strong>of</strong> demethylenation<br />

The methylenedioxyphenyl moiety is found widely distributed in essential oils, alkaloids,<br />

<strong>and</strong> <strong>other</strong> physiologically active compounds <strong>of</strong> natural origin. Compounds containing this<br />

moiety are extensively utilized in food additives, perfumes, medicinals, topical<br />

preparations, <strong>and</strong> insecticides (Fishbein <strong>and</strong> Falk 1969). The methylenedioxybenzenes<br />

(1,3-benzodioxoles) contain a blocked catechol system, which is not considered to be<br />

particularly sensitive to mild hydrolyzing, reducing, or oxidizing reagents (Hennessy<br />

1965). Proposed metabolic pathways for the oxidation <strong>of</strong> methylenedioxyphenyl<br />

compounds <strong>by</strong> P450s <strong>and</strong> APOs are summarized in Figure 42.<br />

Oxidation at the methylene carbon atom (I) causes the generation <strong>of</strong> a putative 2-hydroxy-<br />

1,3-benzodioxole intermediate (II). The intermediate gives rise to a carbene (IV) <strong>by</strong><br />

dehydration or to an o-hydroxyphenyl formate (V; formic acid ester). The base-catalyzed<br />

proton abstraction would yield a carbenium ion (III) that could undergo an α-elimination<br />

107


DISCUSSION<br />

reaction to form also the carbene (Anders et al. 1984). The carbene may either generate<br />

very stable 1,3-benzodioxol-2-carbene complexes (VI) with the heme <strong>of</strong> P450s in the<br />

ferrous state, characterized <strong>by</strong> a Soret peak at 455 nm (Philpot <strong>and</strong> Hodgson 1972), or it is<br />

hydrolyzed to the corresponding formate ester (V) (Murray 2000). Casida <strong>and</strong> co-workers<br />

(Casida et al. 1966) have shown that the release <strong>of</strong> formate- 14 C upon scission <strong>of</strong> the<br />

hydroxy-methylene- 14 C-dioxyphenyl group would lead to both the formation <strong>of</strong> 14 CO 2 ,<br />

<strong>and</strong> also the introduction <strong>of</strong> formic acid- 14 C into the general metabolic pool. The<br />

metabolism <strong>of</strong> 1,3-benzodioxoles to catechols (VII), carbon monoxide <strong>and</strong> formic acid is<br />

consistent with hydroxylation <strong>of</strong> the carbene complex (VI) (Correia et al. 2005,<br />

Simonneaux et al. 2006). During the conversion <strong>of</strong> 1,3-benzodioxoles with APOs,<br />

catechols <strong>and</strong> formic acid were formed in equimolar amounts, i.e. the pathway seemingly<br />

follows the reaction cascade from (I) via (II) <strong>and</strong> (V) to (VII) <strong>and</strong> formic acid.<br />

The stoichiometrical experiments with AaeAPO <strong>and</strong> safrole showed that two hydrogen<br />

peroxide molecules are consumed when one molecule <strong>of</strong> formic acid is formed (Table 20,<br />

Chapter 3.7.2). The same ratio (H 2 O 2 :formic acid, 2:1) was obtained in stoichiometrical<br />

experiments with 4-nitrobenzodioxole, in the case <strong>of</strong> which also the 4-nitrocatechol<br />

formation was followed <strong>by</strong> HPLC; the molecular ratio <strong>of</strong> formic acid:4-nitrocatechol was<br />

1:1 (Appendix, Table 1). This could mean that the conversion <strong>of</strong> 1,3-benzodioxoles to<br />

catechols (VII) <strong>and</strong> formic acid proceeds in two enzymatic steps. A possible explanation<br />

can be suggested (but with the proviso that also <strong>other</strong> reactions are conceivable). APOs<br />

may catalyze the conversion <strong>of</strong> 1,3-benzodioxoles in parallel using two pathways. The first<br />

pathway via the 2-hydroxy-1,3-benzodioxole intermediate (II) may give o-hydroxyphenyl<br />

formate (V), which is hydrolyzed to a catechol (VII) <strong>and</strong> formic acid <strong>and</strong> in the <strong>other</strong><br />

pathway, it is hydroxylated to o-hydroxyphenyl formic acid (X) or the 2,2-dihydroxy-1,3-<br />

benzodioxole intermediate (VIII) (at equilibrium with cyclic phenylene carbonate IX),<br />

which then break down to a catechol (VII) <strong>and</strong> carbon dioxide. This would explain the<br />

double consumption <strong>of</strong> hydrogen peroxide; however, the formation <strong>of</strong> carbon dioxide has<br />

not been proved yet, <strong>and</strong> the observed molar ratio <strong>of</strong> formic acid <strong>and</strong> 4-nitrocatechol <strong>of</strong> 1:1<br />

also contradicts this assumption. Considering the latter fact <strong>and</strong> the presence <strong>of</strong> ascorbic<br />

acid in the reaction mixture, an<strong>other</strong> explanation is possible. Part <strong>of</strong> the peroxide may be<br />

"uselessly" consumed during the peroxidation <strong>of</strong> formed catechols whose phenoxyl<br />

radicals in turn can be re-reduced <strong>by</strong> ascorbic acid (see also Figure 41). Eventually also the<br />

catalase-like side activity <strong>of</strong> APOs that leads to the "unproductive" destruction <strong>of</strong> H 2 O 2<br />

may have to be taken into consideration (R. Ullrich, personal communication).<br />

108


DISCUSSION<br />

Figure 42 Proposed mechanism <strong>of</strong> benzodioxole demethylenation <strong>by</strong> P450 <strong>and</strong> AaeAPO. I-1,3-<br />

benzodioxole; II - 2-hydroxy-1,3-benzodioxole intermediate; III - carbanion; IV - carbene;V - o-<br />

hydroxyphenyl formate; VI - 1,3-benzodioxol-2-carbene complexes; VII - catechol derivative;<br />

VIII - 2,2-dihydroxy-1,3-benzodioxole intermediate; IX - cyclic phenylene carbonate; X - o-<br />

hydroxyphenyl formic acid.<br />

4.3 NBD assay<br />

4-Nitrocatechol is a colorimetric, pH-selective indicator compound showing a yellow color<br />

under acidic conditions <strong>and</strong> a red color under alkaline conditions in aqueous solutions<br />

(Cooper <strong>and</strong> Tulane 1936b). This property has already been in use for a long time, in<br />

various enzyme assays where 4-nitrocatechol is the diagnostic reaction product (e.g. P450<br />

monooxygenase-catalyzed oxidation <strong>of</strong> 4-nitrophenol (Tassaneeyakul et al. 1993a,<br />

Tassaneeyakul et al. 1993b) or sulfatase-catalysed hydrolysis <strong>of</strong> 4-nitrocatechol sulfate<br />

(Baum et al. 1959, Dzialoszyński 1955)). Here, we developed an assay on the basis <strong>of</strong> 4-<br />

nitrocatechol formation from nitrobenzodioxole (NBD), which is suitable to selectively<br />

detect peroxygenase activities in complex growth media with high background absorption<br />

such as soybean-peptone suspensions (Anh et al. 2007, Ullrich et al. 2004). Such growth<br />

media are typically colored (high background absorption), which may interfere with<br />

classic peroxygenase assays that follow substrate oxidation in the UV range.<br />

The exact measurement <strong>of</strong> peroxygenase activities is an important prerequisite for studying<br />

them in the lab, for example, during <strong>fungal</strong> fermentations in bioreactors, heterologous<br />

expression studies, protein purification <strong>and</strong> biochemical enzyme characterization as well as<br />

109


DISCUSSION<br />

during the optimization <strong>of</strong> enzymatic in vitro reactions. All these approaches require<br />

simple <strong>and</strong> rapid enzyme assays, which are selective <strong>and</strong> based on simple<br />

spectrophotometric detection. So far, only two spectrophotometric peroxygenase assays<br />

have been reported: (i) the oxidation <strong>of</strong> veratryl alcohol (3,4-dimethoxybenzyl alcohol) or<br />

methyl 3,4-dimethoxybenzyl ether to veratraldehyde (3,4-dimethoxybenzaldehyde) (Kinne<br />

et al. 2009b, Ullrich et al. 2004), which is followed in the UV range at 310 nm <strong>and</strong> lacks<br />

full selectivity, since there is some overlap with lignin peroxidase activity under acidic<br />

conditions (pH 3-5) (H<strong>of</strong>richter <strong>and</strong> Ullrich 2010, Schmidt et al. 1989, Tien et al. 1986),<br />

<strong>and</strong> (ii) the oxidation <strong>of</strong> naphthalene to 1-naphthol (Kluge et al. 2007), which proceeds in<br />

the UV range as well (at 303 nm). The assay established here is a simple method for the<br />

rapid spectrophotometric detection (in the visible range) <strong>of</strong> <strong>fungal</strong> peroxygenase activity in<br />

vivo <strong>and</strong> in vitro. The NBD assay can be applied for the spectrophotometric identification<br />

<strong>and</strong> quantification <strong>of</strong> <strong>fungal</strong> peroxygenase activities, (i) in complex plant-based media<br />

during <strong>fungal</strong> growth, (ii) in environmental samples, (iii) for reaction control during<br />

enzymatic catalysis <strong>and</strong> (iv) for the detection <strong>of</strong> peroxygenase-positive fungi in the course<br />

<strong>of</strong> high-throughput screenings <strong>of</strong> wild-type <strong>and</strong> recombinant strains as well as mutants.<br />

4.4 Long-term <strong>and</strong> in-vivo experiments<br />

The purification <strong>of</strong> peroxygenases is quite complicated <strong>and</strong> time-consuming, <strong>and</strong> would be<br />

too expensive for most kinds <strong>of</strong> industrial applications (Gröbe et al. 2012, Ullrich et al.<br />

2009). For this reason, it is important to find possibilities to apply the enzymes in a more<br />

economic way, either <strong>by</strong> using crude preparations or extending their "life-span", for<br />

example, <strong>by</strong> immobilization <strong>and</strong>/or re-use over several cycles (see also chapter 3.5). Initial<br />

tests in this direction were performed <strong>by</strong> retaining <strong>and</strong> re-using the enzyme during<br />

propranolol hydroxylation in Vivaspins® (special centrifugal concentrators with<br />

ultrafiltration membranes). Later long-term experiments were carried out in hollow-fiber<br />

modules holding back the enzyme from the circulating filtrate containing the reactants <strong>of</strong><br />

enzymatic conversion. AaeAPO turned out to be very stable under these conditions <strong>and</strong><br />

was successfully used over a period <strong>of</strong> two weeks. Over the whole time, the enzyme was<br />

active <strong>and</strong> continuously hydroxylated the model substrate propranolol into 5-<br />

hydroxypropranolol (see Figure 35, chapter 3.17.2). The cumulative molar amount <strong>of</strong><br />

product (50.4 µmol) <strong>and</strong> the amount <strong>of</strong> applied enzyme (7.9 nmol) show that one<br />

peroxygenase molecule oxidized about 6,400 substrate molecules in the course <strong>of</strong> the<br />

experiment. This is a promising basis for the further improvement <strong>of</strong> HDM production<br />

110


DISCUSSION<br />

with <strong>fungal</strong> peroxygenases. Existing methods use either multi-step organic synthesis<br />

(Johansson et al. 2007, Oatis et al. 1981), whole-cell biotransformations with cell cultures<br />

(e.g. human hepatocytes (Acikgöz et al. 2009, Li et al. 1999, Maurel 1996) or microbial<br />

host cells (e.g. Schizosaccharomyces pombe, Nocardia autotropica or Curvularia lunata;<br />

(Julsing et al. 2008, Peters et al. 2007) as well as, at least at lab scale, engineered P450<br />

enzymes (Leemann et al. 1993, Schroer et al. 2010, Weis et al. 2009).<br />

An<strong>other</strong> promising field <strong>of</strong> peroxygenase application could be the removal <strong>of</strong><br />

<strong>pharmaceuticals</strong> from waste water or <strong>other</strong> environmental media. The occurrence <strong>of</strong> watersoluble<br />

<strong>and</strong> pharmacologically active organic micropollutants in the effluents <strong>of</strong><br />

wastewater treatment plants (WWTPs) has caused much attention over the recent years<br />

(Lienert et al. 2007). It has been a matter <strong>of</strong> concern that the existing WWTPs are<br />

inefficient in removing some <strong>of</strong> these pollutants <strong>and</strong> there is an increasing interest to<br />

develop efficient techniques for achieving their destruction (Clara et al. 2005, Marco-Urrea<br />

et al. 2009). A number <strong>of</strong> experiments with the white-rot fungus Trametes versicolor<br />

showed that waste waters contaminated with several <strong>pharmaceuticals</strong> such as cl<strong>of</strong>ibric<br />

acid, ibupr<strong>of</strong>en <strong>and</strong> carbamazepine could be cleaned to some extent <strong>by</strong> the <strong>fungal</strong><br />

mycelium. However, in these experiments, adsorption <strong>and</strong> transformation <strong>of</strong><br />

<strong>pharmaceuticals</strong> <strong>by</strong> the fungus were not clearly distinguished, nor were intra- <strong>and</strong><br />

extracellular enzymatic mechanisms (Blánquez et al. 2004). The concentrations tested in<br />

this study (~10 mg L -1 ) were also three to five orders <strong>of</strong> magnitude higher than those<br />

usually present in effluents <strong>of</strong> municipal sewage plants (0.1-50-µg L -1 range). Subsequent<br />

studies dealing with the analgesic ketopr<strong>of</strong>en at an environmentally relevant concentration<br />

(40 µg L -1 ) showed that the substrate was almost completely removed <strong>by</strong> T. versicolor<br />

(Marco-Urrea et al. 2010b). Tests with environmentally relevant concentrations <strong>of</strong><br />

propranolol carried out herein demonstrated that AaeAPO is able to remove about 50% <strong>of</strong><br />

the compound at a concentration as low as 1.4 µg L -1 . The experiments were done with the<br />

purified enzyme <strong>and</strong> respective controls, so that all <strong>other</strong> paths <strong>of</strong> substance removal such<br />

as absorption or involvement <strong>of</strong> <strong>other</strong> enzymes than APO could be ruled out. The positive<br />

result obtained, despite the fact that the K m for propranolol is rather high (1,340 µM), can<br />

be explained <strong>by</strong> the reaction cycle <strong>of</strong> APO <strong>and</strong> the particular reaction conditions. Thus, the<br />

continuous dosage <strong>of</strong> hydrogen peroxide with a syringe pump may have ensured the<br />

(pre)formation <strong>of</strong> sufficient amounts <strong>of</strong> a reactive APO species (compound I; (Wang et al.<br />

2012). The presence <strong>of</strong> a certain steady-state concentration <strong>of</strong> compound I (spontaneous<br />

decomposition rate 1.4 s -1 ) combined with a sufficient reaction time (>12 hours) under<br />

111


DISCUSSION<br />

stirring may enable the oxidation <strong>of</strong> quite low amounts <strong>of</strong> the substrate independent on the<br />

high apparent K m value.<br />

In vivo experiments with A. aegerita were carried out to follow the fate <strong>of</strong> two<br />

pharmaceutical model compounds (propranolol, dicl<strong>of</strong>enac) in <strong>fungal</strong> cultures <strong>and</strong> to find<br />

indications for the involvement <strong>of</strong> APOs in biodegradation processes under natural<br />

conditions. The transformation <strong>of</strong> the anti-inflammatory drug dicl<strong>of</strong>enac had already been<br />

studied in liquid cultures <strong>of</strong> Phanerochaete chrysosporium under fed-batch conditions. The<br />

fungus started to eliminate dicl<strong>of</strong>enac (34 µM) already after two hours (77% removal) <strong>and</strong><br />

after 18 hours, 94% <strong>of</strong> the compound were converted (Rodarte-Morales et al. 2012). The<br />

capability <strong>of</strong> dicl<strong>of</strong>enac degradation <strong>by</strong> T. versicolor was demonstrated <strong>by</strong> Marco-Urrea et<br />

al. (Marco-Urrea et al. 2010a). They observed an almost complete dicl<strong>of</strong>enac removal<br />

(>94%) within the first hour <strong>of</strong> <strong>fungal</strong> treatment in the presence <strong>of</strong> <strong>fungal</strong> pellets; the drug<br />

was added both at relatively high (10 mg L -1 ) <strong>and</strong> low, environmentally relevant (45 µg L -<br />

1 ) concentrations in a defined liquid medium. During this treatment, 4'-hydroxydicl<strong>of</strong>enac<br />

<strong>and</strong> 5-hydroxydicl<strong>of</strong>enac were detected as metabolites, which disappeared within the next<br />

24 hours (Cruz-Morató et al. 2012, Marco-Urrea et al. 2010a). An<strong>other</strong> example <strong>of</strong> using<br />

the <strong>fungal</strong> mycelium as biodegradative agent is the so-called biologically advanced<br />

oxidation <strong>of</strong> propranolol <strong>by</strong> T. versicolor (80% removal within 6 hours <strong>of</strong> incubation). In<br />

this case, extracellular oxidizing species (e.g. hydroxyl radicals) were thought to be<br />

produced through a quinone redox cycling mechanism catalyzed <strong>by</strong> an intracellular<br />

quinone reductase <strong>and</strong> ligninolytic enzymes <strong>of</strong> Trametes versicolor, after addition <strong>of</strong><br />

lignin-derived 2,6-dimethoxy-1,4-benzoquinone <strong>and</strong> Fe 3+ -oxalate (Marco-Urrea et al.<br />

2010c).<br />

Dicl<strong>of</strong>enac <strong>and</strong> propranolol also readily disappeared in liquid cultures <strong>of</strong> A. aegerita,<br />

interestingly when a high peroxygenase level (ca. 600 U L -1 ) was present. The<br />

experiments, carried out in parallel in two different growth media (one <strong>of</strong> which stimulated<br />

peroxygenase production (Ullrich et al. 2004)) indeed indicated an involvement <strong>of</strong><br />

peroxygenase in this process. However, no transformation products were detectable <strong>by</strong><br />

HPLC, probably due to coupling/polymerization reactions, which could not be prevented<br />

in contrast to in vitro experiments where ascorbic acid served as radical scavenger.<br />

Furthermore, analyses <strong>of</strong> control cultures <strong>of</strong> A. aegerita (soybean medium) showed the<br />

presence <strong>of</strong> numerous aromatic ingredients, which disappeared in the course <strong>of</strong> <strong>fungal</strong><br />

growth. UV-Vis spectra <strong>of</strong> these substances indicated a polyphenolic nature with<br />

similarities to flavonoids (Figure 30) (Barková et al. 2011). One-electron oxidation <strong>of</strong><br />

112


DISCUSSION<br />

these compounds would yield phenoxyl radicals, which may force the overall<br />

polymerization process. The peroxidative activity <strong>of</strong> AaeAPO, or the activity <strong>of</strong> laccase<br />

which is also secreted <strong>by</strong> A. aegerita (Ullrich et al. 2004), may be responsible for these<br />

reactions. Interestingly, in both growth media (soybean medium, Czapek-Dox) the color<br />

changed to violet when dicl<strong>of</strong>enac was added on day six. The same color was observed in<br />

in vitro studies with APOs in the absence <strong>of</strong> ascorbic acid (Appendix Figure 2). On the<br />

basis <strong>of</strong> UV-Vis <strong>and</strong> mass spectra as well as literature data from Miyamoto at al. (1997)<br />

we propose that this compound is the one-electron-oxidation product dicl<strong>of</strong>enac-1',4'-<br />

quinone imine (Figure 43) (Miyamoto et al. 1997, Waldon et al. 2010).<br />

Figure 43 Proposed mechanism <strong>of</strong> dicl<strong>of</strong>enac-1',4'-quinone imine formation <strong>by</strong> AaeAPO in vitro<br />

without ascorbic acid <strong>and</strong> <strong>by</strong> <strong>by</strong> <strong>fungal</strong> cultures. In Czapek-Dox medium, conversion may be only<br />

possible <strong>by</strong> intracellular 450s (no peroxygenase formation; laccase alone is not able to catalyze this<br />

reaction) (Marco-Urrea et al. 2010a). In soybean medium, this pathway could be realized both <strong>by</strong><br />

intracellular P450s or secreted AaeAPO or a mixed AaeAPO-laccase system (Figure 48;<br />

(H<strong>of</strong>richter et al. 2010, Ullrich <strong>and</strong> H<strong>of</strong>richter 2007, Waldon et al. 2010).<br />

4.5 Homology model structures <strong>of</strong> APOs.<br />

In order to better underst<strong>and</strong> the catalytic properties <strong>of</strong> APOs <strong>and</strong> differences between<br />

them, homology modelling was performed. The full-length sequences <strong>of</strong> the<br />

peroxygenases from A. aegerita, C. radians <strong>and</strong> M. rotula have recently been obtained <strong>and</strong><br />

confirm the enzymes’ affiliation to the heme-thiolate proteins. The sequences revealed no<br />

homology to classic heme peroxidases or cytochrome P450 enzymes, <strong>and</strong> only little<br />

113


DISCUSSION<br />

114<br />

DISCUSSION<br />

Figure 45 Homology model structures <strong>of</strong> CraAPO (left), AaeAPO (middle) <strong>and</strong> MroAPO (right). Homology modelling was performed with I-TASSER<br />

s<strong>of</strong>tware (Roy et al. 2010) using the crystal structure <strong>of</strong> unspecific peroxygenase <strong>of</strong> A. aegerita as basis [(Piontek et al. 2010), as well as unpublished data, the<br />

same authors]. The resulting models <strong>of</strong> CraAPO <strong>and</strong> MroAPO were superimposed with the AaeAPO structure using the PyMOL program (2010), resulting in<br />

a root-mean-square deviation (RMS) <strong>of</strong> 1.133 over 295 aligned residues for CraAPO, <strong>and</strong> a RMS <strong>of</strong> 1.381 over 177 aligned residues for MroAPO. Prepared<br />

<strong>by</strong> Marek Pecyna <strong>and</strong> Marzena Poraj-Kobielska (2012).


DISCUSSION<br />

homology (


DISCUSSION<br />

For comparison, the structure <strong>of</strong> propranolol, which is smaller <strong>and</strong> a good substrate <strong>of</strong> the<br />

AaeAPO, was placed into the active site <strong>and</strong> fit it well (Figure 42, examples <strong>of</strong> <strong>other</strong><br />

structures are given in the Appendix). The homology model <strong>of</strong> CraAPO shown in Figure<br />

45 indicates that the entrance to the heme channel is about the size <strong>of</strong> that <strong>of</strong> AaeAPO <strong>and</strong><br />

the active site volume is even larger, but one phenylalanine seems to block the entrance<br />

(Figure 49).<br />

The position at the end <strong>of</strong> the polypeptide chain (C-terminus) indicates that this<br />

phenylalanine may not be fixed <strong>and</strong> can move to some extent to open <strong>and</strong> close the<br />

entrance <strong>of</strong> the heme channel, i.e. it may act as a sort <strong>of</strong> molecular "gatekeeper". This may<br />

explain the overall lower conversion rates observed for CraAPO compared to AaeAPO, in<br />

spite <strong>of</strong> a similar amino acids sequence <strong>and</strong> configuration (59% sequence identity <strong>and</strong> 69%<br />

similarity; for comparison, MraAPO shares only about 30% sequence identity with both<br />

enzymes) (Pecyna et al. 2009), personal communication). Of course, more precise<br />

predictions will only be possible when the crystal structures <strong>of</strong> all three APOs will have<br />

been published.<br />

Figure 48 Amino acids that possess reactive groups <strong>and</strong> are found in the active sites <strong>of</strong> AaeAPO,<br />

CraAPO <strong>and</strong> MroAPO. Tyr - tyrosine, Phe - phenylalanine, Glu – glutamic acid, Thr – threonine,<br />

Ser – serine <strong>and</strong> Arg - arginine.<br />

Figure 49 shows homology model structures <strong>of</strong> the active site <strong>of</strong> the three APOs studied<br />

here.<br />

The channels <strong>of</strong> AaeAPO <strong>and</strong> CraAPO are rich in aromatic amino acids, mainly<br />

phenylalanines (Phe) <strong>and</strong> a tyrosine (Tyr), whereas in MroAPO, almost exclusively<br />

aliphatic amino acids are found (Ile, Leu, Val, Ala). Phe is a hydrophobic aromatic amino<br />

acid with a benzyl side chain; Tyr possesses also an aromatic ring that is however less<br />

hydrophobic than that <strong>of</strong> Phe due to its phenolic OH-group.<br />

116


DISCUSSION<br />

117<br />

Figure 46 Homology models <strong>of</strong> peroxygenase channels (solvent accessible surface calculation) leading to the active sites <strong>of</strong> AaeAPO (B) <strong>and</strong> MroAPO (C)<br />

fitting codeine. Dimensions (in Ǻ) <strong>of</strong> the heme channels are shown in (A; AaeAPO) <strong>and</strong> (F; MroAPO). A 3-Dimensional model <strong>of</strong> codeine is given in D1, D2<br />

<strong>and</strong> D3 from different perspectives. (E) illustrates the active site's architecture <strong>of</strong> MroAPO with codeine as ball-<strong>and</strong>-stick model: heme (red), magnesium ion<br />

(green), codeine substrate (violet), amino acids (gray).


DISCUSSION<br />

118<br />

DISCUSSION<br />

Figure 47 Homology models <strong>of</strong> heme channels with active sites fitting hydrocortisone. AaeAPO from the top (A) <strong>and</strong> from the side (B) with channel length in<br />

Ǻ <strong>and</strong> MroAPO from the top (F) <strong>and</strong> side (E). (C) Electron surface density <strong>of</strong> hydrocortisone with the length <strong>of</strong> the structure, <strong>and</strong> (D) stick model <strong>of</strong><br />

hydrocortisone from two <strong>other</strong> perspectives (dimensions in Ǻ).


DISCUSSION<br />

Figure 49 Heme channels leading to the active site <strong>of</strong> AaeAPO, CraAPO <strong>and</strong> MroAPO, including<br />

important amino acid residues. Aromatic amino acids (orange): Phe - phenylalanine, Tyr - tyrosine;<br />

functional amino acids (blue): Glu - glutamic acid, Ser – serine, Arg - arginine, Thr - threonine;<br />

aliphatic amino acids (grey): Ile - isoleucine, Leu - leucine, Ala - alanine, Gly - glycine, Val -<br />

valine <strong>and</strong> Pro - proline. One example <strong>of</strong> each kind <strong>of</strong> amino acid is labeled with the respective<br />

abbreviation (model <strong>and</strong> figure prepared <strong>by</strong> M. Pecyna <strong>and</strong> M. Poraj-Kobielska, 2012).<br />

119


DISCUSSION<br />

Near the heme <strong>and</strong> the stabilizing magnesium, three aliphatic amino acids are found:<br />

glutamic acid (Glu), arginine (Arg) <strong>and</strong> threonine (Thr; Figure 48)(Ophardt 2003). Glu has<br />

a carboxylic side chain <strong>and</strong> can exist as its negatively charged deprotonated carboxylate<br />

(Pecyna et al. 2009, Piontek et al. 2010, Sundaramoorthy et al. 1995). As an acid-base<br />

catalyst, it may play a key role in peroxide binding <strong>and</strong> cleavage, <strong>and</strong> there<strong>by</strong> in the<br />

formation <strong>of</strong> AaeAPO compound I (Wang et al. 2012). Arg may act as a charge stabilizer<br />

in this process as in CPO (Sundaramoorthy 1995) <strong>and</strong> Thr is probably the proton acceptor.<br />

In MroAPO, serine (Ser) seemingly replaces Arg. In classical peroxidases, Glu is<br />

substituted <strong>by</strong> a histidine (His), whereas in DyP-type peroxidases an aspartic acid (Asp)<br />

replaces Glu while the charge stabilizer is again an Arg (Sugano et al. 2007).<br />

4.6 Key findings<br />

1. All tested <strong>fungal</strong> unspecific/aromatic peroxygenases (APOs from Agrocybe<br />

aegerita, Coprinellus radians, Marasmius rotula) catalyzed the<br />

oxygenation/hydroxylation or N-/O-dealkylation (Figure 50) <strong>of</strong> diverse<br />

<strong>pharmaceuticals</strong> <strong>and</strong> illicit <strong>drugs</strong>, in most cases with high regioselectivity. Many <strong>of</strong><br />

the oxidation products were identical to human drug metabolites formed <strong>by</strong> P450s.<br />

2. The peroxygenase <strong>of</strong> M. rotula (MroAPO) was moreover found to catalyze the<br />

conversion <strong>of</strong> different steroids with preference for those bearing an oxohydroxyethyl<br />

side chain.<br />

3. The different substrate specificities <strong>and</strong> selectivities <strong>of</strong> individual APOs raise the<br />

possibility to choose the appropriate enzyme according to the needed product<br />

("catalytic toolbox" for oxyfunctionalizations).<br />

4. The AaeAPO-catalyzed oxygenation <strong>of</strong> aromatic <strong>pharmaceuticals</strong> such as<br />

propranolol, dicl<strong>of</strong>enac or carbamazepine <strong>and</strong> the hydroxylation <strong>of</strong> benzylic<br />

carbons in compounds such as tolbutamide resulted in the incorporation <strong>of</strong> an 18 O-<br />

atom from H 18 2 O 2 into the reaction products (phenols, alcohols), which identifies<br />

these reactions as true peroxygenations.<br />

5. The stoichiometry <strong>of</strong> AaeAPO-catalyzed sildenafil oxidation showed that the<br />

reaction was a two-electron oxidation yielding an N-dealkylated product (Ndesmethylsildenafil)<br />

<strong>and</strong> an aldehyde (formaldehyde).<br />

6. The deethylenation <strong>of</strong> phenacetin-d1 <strong>by</strong> AaeAPO showed an observed<br />

intramolecular deuterium isotope effect (k H /k D ) obs <strong>of</strong> 3.1, which is close to the<br />

values that have been observed for the P450-catalyzed O-dealkylation <strong>of</strong> this<br />

120


DISCUSSION<br />

substrate. This indicates that phenacetin oxidation <strong>by</strong> APOs may proceed via an<br />

electron transfer rather than a hydrogen abstraction mechanism.<br />

7. The kinetic data <strong>and</strong> particularly the catalytic turnover numbers (k cat ) determined<br />

for the conversion <strong>of</strong> various <strong>pharmaceuticals</strong> <strong>by</strong> AaeAPO suggest that APOs may<br />

be at least as efficient as the best P450s <strong>and</strong> hence represent useful alternatives for<br />

the regioselective preparation <strong>of</strong> human drug metabolites (HDMs).<br />

8. Steady-state bisubstrate kinetics <strong>of</strong> the AaeAPO-catalyzed demethylenation <strong>of</strong><br />

nitro-1,3-benzodioxole (NBD), which yielded catechol <strong>and</strong> formic acid, gave<br />

parallel double reciprocal plots suggestive <strong>of</strong> a ping-pong mechanism.<br />

9. The NBD assay established on the basis <strong>of</strong> this reaction is a simple method for the<br />

rapid spectrophotometric detection <strong>of</strong> <strong>fungal</strong> peroxygenase activities in vivo <strong>and</strong> in<br />

vitro.<br />

10. The testing <strong>of</strong> twelve validated P450 assays with the three APOs showed that they<br />

possess a substrate spectrum <strong>and</strong> an activity range that overlap with those <strong>of</strong> the<br />

most important human P450s involved in drug metabolism.<br />

11. Stoichiometry analysis <strong>of</strong> AaeAPO-catalyzed safrole oxidation showed that two<br />

hydrogen peroxide molecules were consumed during the formation <strong>of</strong> one molecule<br />

<strong>of</strong> products (catechol <strong>and</strong> formic acid). This indicates a divergent mechanism that<br />

may involve two different pathways <strong>and</strong>/or unproductive peroxide-consuming side<br />

reactions.<br />

12. APOs are able to use, in addition to H 2 O 2 , different organic peroxides as cooxidant,<br />

<strong>and</strong> they efficiently catalyze oxidation reactions in the presence <strong>of</strong><br />

different peroxide-generating systems.<br />

13. AaeAPO was capable <strong>of</strong> oxidizing the pharmaceutical propranolol at a very low,<br />

environmentally relevant concentration (in the ppb range).<br />

14. APOs could be used in long-term in-vitro experiments over several days for the<br />

constant synthesis <strong>of</strong> the HDM, 5-hydroxypropranolol.<br />

15. Propranolol <strong>and</strong> dicl<strong>of</strong>enac disappeared rapidly when added to growing cultures <strong>of</strong><br />

A. aegerita indicating <strong>fungal</strong> drug degradation also under natural conditions.<br />

16. Homology models <strong>of</strong> APOs provide explanations for limitations observed with<br />

regard to the substrate size <strong>and</strong> structure. The size <strong>of</strong> the entrance to the heme<br />

channel, the amino acids there <strong>and</strong> the active site volume seem to be key<br />

parameters determing peroxygenase performance.<br />

121


DISCUSSION<br />

122<br />

DISCUSSION<br />

Figure 50 Mechanistic routes <strong>of</strong> reactions catalyzed <strong>by</strong> <strong>fungal</strong> peroxygenases with <strong>pharmaceuticals</strong> <strong>and</strong> <strong>drugs</strong>. From left to right: demethylenation (ecstasy),<br />

N-dealkylation (lidocaine), O-dealkylation (naproxen), aliphatic side chain hydroxylation (tolbutamide), aromatic epoxidation/hydroxylation (dicl<strong>of</strong>enac),<br />

phenol oxidation (morphine); modified according to (H<strong>of</strong>richter et al. 2010)


DISCUSSION<br />

The most important finding <strong>of</strong> this work is that APOs are capable <strong>of</strong> oxidizing a large<br />

variety <strong>of</strong> <strong>pharmaceuticals</strong> <strong>and</strong> <strong>other</strong> <strong>drugs</strong>, <strong>of</strong>ten yielding typical human drug metabolites.<br />

This opens the fascinating possibility to produce diverse human drug metabolites in the<br />

future using a "peroxygenase toolbox", figuratively speaking as an artificial liver in the<br />

reaction tube.<br />

4.7 Outlook<br />

In the meantime, one <strong>of</strong> the most promising approaches in peroxygenase research, the<br />

empirical comparison <strong>of</strong> as many APOs as possible to obtain a broad selection <strong>of</strong><br />

biocatalysts with different substrate specificities, has taken shape. Thus, recent genetic<br />

investigations have resulted in a phylogenic tree, which impressively shows the widespread<br />

occurrence <strong>of</strong> peroxygenases in the whole <strong>fungal</strong> kingdom (Pecyna et al. 2009). New wildtype<br />

peroxygenases have been found in a number <strong>of</strong> basidiomycetes, for example, in<br />

Coprinopsis verticillata (Anh et al. 2007), Agrocybe parasitica <strong>and</strong> Auricularia auriculajudae<br />

(R. Ullrich 2012, personal communication). Furthermore, the first attempts to<br />

heterologously produce peroxygenases in suitable hosts (e.g. Aspergillus spp.,<br />

Saccharomyces cerevisiae) have been successful (H. Lund, Novozymes A/S & M. Alcalde,<br />

CSIC Madrid, 2012; personal communications). The above developments open the<br />

possibility to extend the APO-based “monooxygenation toolbox” for the production <strong>of</strong><br />

<strong>other</strong> HDMs <strong>and</strong> fine chemicals. Also the preparation <strong>of</strong> HDMs <strong>and</strong> fine chemicals even at<br />

industrial scale seems to be reasonable taking into account the positive long-term<br />

experiments presented here <strong>and</strong> assuming that the heterologous mass production <strong>of</strong><br />

peroxygenases will be successful in the near future. Environmental studies performed over<br />

the last years indicate that there is a strong need to remove <strong>pharmaceuticals</strong> <strong>and</strong> their<br />

metabolites from the environment. In this context, peroxygenases could become an<br />

innovative biocatalytic tool to reduce the risk potential <strong>of</strong> <strong>pharmaceuticals</strong>. The properties<br />

<strong>of</strong> peroxygenases such as: i) extracellular location, ii) active secretion <strong>by</strong> the fungi, iii)<br />

high stability <strong>and</strong> iv) independence from complex, expensive c<strong>of</strong>actors, as well as the<br />

positive results in the removal <strong>of</strong> low-concentration <strong>pharmaceuticals</strong> make the application<br />

<strong>of</strong> APOs even in the water-treatment sector conceivable. Again, the heterologous largescale<br />

production <strong>of</strong> peroxygenases, i.e. their availability in tons per year, will be de rigueur<br />

to achieve this challenging goal.<br />

123


DISCUSSION<br />

Figure 51 Phylogenetic tree <strong>of</strong> peroxygenases (unpublished results, Kellner 2012).<br />

124


REFERENCES<br />

5 References<br />

Acikgöz, A., Karim, N., Giri, S., Schmidt-Heck, W. <strong>and</strong> Bader, A. (2009), Two<br />

compartment model <strong>of</strong> diazepam biotransformation in an organotypical culture <strong>of</strong><br />

primary human hepatocytes, Toxicol Appl Pharmacol 234, 179-91.<br />

Alex<strong>and</strong>er, N. J., Hohn, T. M. <strong>and</strong> McCormick, S. P. (1998), The TRI11 Gene <strong>of</strong> Fusarium<br />

sporotrichioides Encodes a Cytochrome P-450 Monooxygenase Required for C-15<br />

Hydroxylation in Trichothecene Biosynthesis, Appl Environ Microb 64, 221-25.<br />

Alkatheeri, Wasfi <strong>and</strong> Lambert (1999), Pharmacokinetics <strong>and</strong> metabolism <strong>of</strong> ketopr<strong>of</strong>en<br />

after intravenous <strong>and</strong> intramuscular administration in camels, J Vet Pharmacol<br />

Ther 22, 127-35.<br />

Allain, E. J., Hager, L. P., Deng, L. <strong>and</strong> Jacobsen, E. N. (1993), Highly enantioselective<br />

epoxidation <strong>of</strong> disubstituted alkenes with hydrogen peroxide catalyzed <strong>by</strong><br />

chloroperoxidase, J Am Chem Soc 115, 4415-16.<br />

Anders, M. W., Sunram, J. M. <strong>and</strong> Wilkinson, C. F. (1984), Mechanism <strong>of</strong> the metabolism<br />

<strong>of</strong> 1,3-benzodioxoles to carbon monoxide, Biochem Pharmacol 33, 577-80.<br />

Andersson, T., Miners, J. O., Veronese, M. E. <strong>and</strong> Birkett, D. J. (1994), Diazepam<br />

metabolism <strong>by</strong> human liver microsomes is mediated <strong>by</strong> both S-mephenytoin<br />

hydroxylase <strong>and</strong> CYP3A is<strong>of</strong>orms., Br J Clin Pharmacol 38, 131-37.<br />

Andersson, T., Miners, J. O., Veronese, M. E., Tassaneeyakul, W., Tassaneeyakul, W.,<br />

Meyer, U. A. <strong>and</strong> Birkett, D. J. (1993), Identification <strong>of</strong> human liver cytochrome<br />

P450 is<strong>of</strong>orms mediating omeprazole metabolism, Br J Clin Pharmacol 36, 521-30.<br />

Anh, D. H. (2008), Enzymatic <strong>and</strong> molecular characterization <strong>of</strong> haloperoxidaseperoxygenase<br />

from litter- <strong>and</strong> dung-dwelling agaric fungi, International graduate<br />

school <strong>Zittau</strong>, <strong>Zittau</strong><br />

Anh, D. H., Ullrich, R., Benndorf, D., Svatos, A., Muck, A. <strong>and</strong> H<strong>of</strong>richter, M. (2007), The<br />

coprophilous mushroom Coprinus radians secretes a haloperoxidase that catalyzes<br />

aromatic peroxygenation, Appl Environ Microbiol 73, 5477-85.<br />

Ankley, G. T., Brooks, B. W., Huggett, D. B. <strong>and</strong> Sumpter, a. J. P. (2007), Repeating<br />

History: Pharmaceuticals in the Environment, Environ Sci Technol 41, 8211-17.<br />

Anzenbacher, P. <strong>and</strong> Anzenbacherová, E. (2001), Cytochromes P450 <strong>and</strong> metabolism <strong>of</strong><br />

xenobiotics, Cell Mol Life Sci 58, 737-47.<br />

Apple<strong>by</strong>, A. C. (1967), A soluble haemoprotein P 450 from nitrogen-fixing Rhizobium<br />

bacteroides, BIOCHIM BIOPHYS ACTA (2), 399-402.<br />

Ar<strong>and</strong>a, E., Kinne, M., Kluge, M., Ullrich, R. <strong>and</strong> H<strong>of</strong>richter, M. (2009a), <strong>Conversion</strong> <strong>of</strong><br />

dibenzothiophene <strong>by</strong> the mushrooms Agrocybe aegerita <strong>and</strong> Coprinellus radians<br />

<strong>and</strong> their extracellular peroxygenases, Appl Microbiol Biot 82, 1057-66.<br />

Ar<strong>and</strong>a, E., Ullrich, R. <strong>and</strong> H<strong>of</strong>richter, M. (2009b), <strong>Conversion</strong> <strong>of</strong> polycyclic aromatic<br />

hydrocarbons, methyl naphthalenes <strong>and</strong> dibenz<strong>of</strong>uran <strong>by</strong> two <strong>fungal</strong><br />

peroxygenases, Biodegradation 21, 267-81.<br />

Baillie, T. A. (2009), Approaches to the assessment <strong>of</strong> stable <strong>and</strong> chemically reactive drug<br />

metabolites in early clinical trials, CHEM RES TOXICOL 22, 263-66.<br />

Baillie, T. A.et al. (2002), Drug metabolites in safety testing, Toxicol Appl Pharmacol 182,<br />

188-96.<br />

Ballard, S. A., Gingell, C. J., Tang, K., Turner, L. A., Price, M. E. <strong>and</strong> Naylor, A. M.<br />

(1998), Effects <strong>of</strong> sildenafil on the relaxation <strong>of</strong> human corpus cavernosum tissue in<br />

vitro <strong>and</strong> on the activities <strong>of</strong> cyclic nucleotide phosphodiesterase isozymes, J Urol<br />

159, 2164-71.<br />

Banerjee, N. C., Miller, G. E. <strong>and</strong> Stowe, C. M. (1967), Excretion <strong>of</strong> aminopyrine <strong>and</strong> its<br />

metabolites into cows' milk, Toxicol. Appl. Pharmacol. 10, 604-12.<br />

125


126<br />

REFERENCES<br />

Barková, K., Kinne, M., Ullrich, R., Hennig, L., Fuchs, A. <strong>and</strong> H<strong>of</strong>richter, M. (2011),<br />

Regioselective hydroxylation <strong>of</strong> diverse flavonoids <strong>by</strong> an aromatic peroxygenase,<br />

Tetrahedron 67, 4874-78.<br />

Baum, H., Dodgson, K. <strong>and</strong> Spencer, B. (1959), The assay <strong>of</strong> arylsulphatases A <strong>and</strong> B in<br />

human urine, Clin. Chim. Acta. 4, 453-55.<br />

Beausse, J. (2004), Selected <strong>drugs</strong> in solid matrices: a review <strong>of</strong> environmental<br />

determination, occurrence <strong>and</strong> properties <strong>of</strong> principal substances, TrAC-Trend<br />

Anal Chem 23, 753-61.<br />

Bergé, J., Feyereisen, R. <strong>and</strong> Amichot, M. (1998), Cytochrome P450 monooxygenases <strong>and</strong><br />

insecticide resistance in insects, Philos T Roy Soc B 353, 1701-05.<br />

Bernhardt, R. (2006), Cytochromes P450 as versatile biocatalysts, J BIOTECHNOL 124,<br />

128-45.<br />

Bezalel, L., Hadar, Y., Fu, P. P., Freeman, J. P. <strong>and</strong> Cerniglia, C. E. (1996), Initial<br />

oxidation products in the metabolism <strong>of</strong> pyrene, anthracene, fluorene, <strong>and</strong><br />

dibenzothiophene <strong>by</strong> the white rot fungus Pleurotus ostreatus, Appl Environ<br />

Microbiol 62, 2554-9.<br />

Bisswanger, H. (2000), Enzymkinetik Theorie und Methoden, Wiley-VCH Verlag,<br />

Weinheim.<br />

Bl<strong>and</strong>, T. M., Haining, R. L., Tracy, T. S. <strong>and</strong> Callery, P. S. (2005), CYP2C-catalyzed<br />

delta(9)-tetrahydrocannabinol metabolism: Kinetics, pharmacogenetics <strong>and</strong><br />

interaction with phenytoin, Biochem Pharmacol 70, 1096-103.<br />

Blánquez, P., Casas, N., Font, X., Gabarrell, X., Sarrá , M., Caminal, G. <strong>and</strong> Vicent, T.<br />

(2004), Mechanism <strong>of</strong> textile metal dye biotransformation <strong>by</strong> Trametes versicolor,<br />

Water Res 38, 2166-72.<br />

Bolwell, G. P., Bozak, K. <strong>and</strong> Zimmerlin, A. (1994), Plant cytochrome p450,<br />

Phytochemistry 37, 1491-506.<br />

Boolell, M., Allen, M. J., Ballard, S. A., Gepi-Attee, S., Muirhead, G. J., Naylor, A. M.,<br />

Osterloh, I. H. <strong>and</strong> Gingell, C. (1996), Sildenafil: an orally active type 5 cyclic<br />

GMP-specific phosphodiesterase inhibitor for the treatment <strong>of</strong> penile erectile<br />

dysfunction, Internat Journal Impot Res 8, 47-52.<br />

Booth, K. S., Kimura, S., Lee, H. C., Ikeda-Saito, M. <strong>and</strong> Caughey, W. S. (1989), Bovine<br />

myeloperoxidase <strong>and</strong> lactoperoxidase each contain a high affinity site for calcium,<br />

Biochem Biophys Res Co 160, 897-902.<br />

Bort, R., Macé, K., Boobis, A., Gómez-Lechón, M.-J., Pfeifer, A. <strong>and</strong> Castell, J. (1999),<br />

Hepatic metabolism <strong>of</strong> dicl<strong>of</strong>enac: role <strong>of</strong> human CYP in the minor oxidative<br />

pathways, Biochem. Pharmacol. 58, 787-96.<br />

Bossche, H. V. <strong>and</strong> Koymans, L. (1998), Review Article Cytochromes P450 in fungi,<br />

Mycoses 41, 32-38.<br />

Boxall, A. B. A., Sinclair, C. J., Fenner, K., Kolpin, D. <strong>and</strong> Maund, S. J. (2004), Peer<br />

reviewed: when synthetic chemicals degrade in the environment, Environ Sci<br />

Technol 38, 368A-75A.<br />

Breskvar, K., Ferenčak, Z. <strong>and</strong> Hudnik-Plevnik, T. (1995), The role <strong>of</strong> cytochrome<br />

P45011α in detoxification <strong>of</strong> steroids in the filamentous fungus Rhizopus nigricans,<br />

J Steroid Biochem Mol Bio 52, 271-75.<br />

Brodie, B. B. <strong>and</strong> Axelrod, J. (1949), The fate <strong>of</strong> acetophenetidin (phenacetin) in man <strong>and</strong><br />

methods for the estimation <strong>of</strong> acetophenetidin <strong>and</strong> its metabolites in biological<br />

material, J. Pharmacol. Exp. Ther. 97, 58-67.<br />

Brooks, B. W., Chambliss, C. K., Stanley, J. K., Ramirez, A., Banks, K. E., Johnson, R. D.<br />

<strong>and</strong> Lewis, R. J. (2005), Determination <strong>of</strong> select antidepressants in fish from an<br />

effluent-dominated stream, Environ Toxicol Chem 24, 464-69.<br />

Caputi, L., Di Tullio, A., Di Le<strong>and</strong>ro, L., De Angelis, F. <strong>and</strong> Malatesta, F. (2005), A new<br />

microperoxidase from Marinobacter hydrocarbonoclasticus, Biochim. Biophys.<br />

Acta 1725, 71-80.


REFERENCES<br />

Cardini, G. <strong>and</strong> Jurtshuk, P. (1968), Cytochrome P-450 involvement in the oxidation <strong>of</strong> n-<br />

octane <strong>by</strong> cell-free extracts <strong>of</strong> corynebacterium sp. strain 7E1C, J BIOL CHEM<br />

243, 6070-72.<br />

Carrara, C., Ptacek, C. J., Robertson, W. D., Blowes, D. W., Moncur, M. C., Sverko, E.<br />

<strong>and</strong> Backus, S. (2008), Fate <strong>of</strong> Pharmaceutical <strong>and</strong> Trace Organic Compounds in<br />

Three Septic System Plumes, Ontario, Canada, Environ Sci Technol 42, 2805-11.<br />

Casida, J. E., Engel, J. L., Essac, E. G., Kamienski, F. X. <strong>and</strong> Kuwatsuka, S. (1966),<br />

Methylene-C 14 -Dioxyphenyl Compounds: Metabolism in Relation to Their<br />

Synergistic Action, Science 153, 1130-33.<br />

Celiz, M. D., Tso, J. <strong>and</strong> Aga, D. S. (2009), Pharmaceutical metabolites in the<br />

environment: Analytical challenges <strong>and</strong> ecological risks, Environ Toxicol Chem<br />

28, 2473-84.<br />

Center for Drug Evaluation <strong>and</strong> Research (HFD-130) <strong>and</strong> Food <strong>and</strong> Drug Administration<br />

(2008), Guidance for Industry on Safety Testing <strong>of</strong> Drug Metabolites; Availability,<br />

Food Drug Admin 1-11.<br />

Chapple, C. (1998), Molecular-genetic analysis <strong>of</strong> plant cytochrome P450-dependent<br />

monooxygenases, Annu Rev Plant Phys 49, 311-43.<br />

Chen, Z. R., Somogyi, A. A., Reynolds, G. <strong>and</strong> Bochner, F. (1991), Disposition <strong>and</strong><br />

metabolism <strong>of</strong> codeine after single <strong>and</strong> chronic doses in one poor <strong>and</strong> seven<br />

extensive metabolisers., Br J Clin Pharmacol 31, 381–90.<br />

Cirino, P. C. <strong>and</strong> Arnold, F. H. (2003), Exploring the diversity <strong>of</strong> heme enzymes through<br />

directed evolution, in: Directed molecular evolution <strong>of</strong> proteins, Vol. 215-43 (Dr.<br />

Susanne Brakmann, P. D. K. J., Ed.)<br />

Clara, M., Strenn, B., Gans, O., Martinez, E., Kreuzinger, N. <strong>and</strong> Kroiss, H. (2005),<br />

Removal <strong>of</strong> selected <strong>pharmaceuticals</strong>, fragrances <strong>and</strong> endocrine disrupting<br />

compounds in a membrane bioreactor <strong>and</strong> conventional wastewater treatment<br />

plants, Water Res 39, 4797-807.<br />

Clel<strong>and</strong>, W. W. (1963), The kinetics <strong>of</strong> enzyme-catalyzed reactions with two or more<br />

substrates or products: III. Prediction <strong>of</strong> initial velocity <strong>and</strong> inhibition patterns <strong>by</strong><br />

inspection, Biochim Biophys A Spec Sec Enzymol Sub 67, 188-96.<br />

Clouse, S. D. <strong>and</strong> Sasse, J. M. (1998), Brassinosteroids: essential regulators <strong>of</strong> plant<br />

growth <strong>and</strong> development, Annu Rev Plant Phys 49, 427-51.<br />

Colonna, S., Gaggero, N., Manfredi, A., Casella, L., Gullotti, M., Carrea, G. <strong>and</strong> Pasta, P.<br />

(1990), Enantioselective oxidations <strong>of</strong> sulfides catalyzed <strong>by</strong> chloroperoxidase,<br />

Biochemistry 29, 10465-68.<br />

Cooper, S. R. <strong>and</strong> Tulane, V. J. (1936a), Action <strong>of</strong> 4-Nitrocatechol as a Titration Indicator,<br />

Ind Eng Chem Anal Ed 8, 210-11.<br />

Cooper, S. R. <strong>and</strong> Tulane, V. J. (1936b), Action <strong>of</strong> 4-Nitrocatechol as a Titration Indicator,<br />

Industrial & Engineering Chemistry Analytical Edition 8, 210-11.<br />

Correia, M., Ortiz de Montellano, P. <strong>and</strong> Ortiz de Montellano, P. R. (2005), Inhibition <strong>of</strong><br />

Cytochrome P450 Enzymes<br />

Cytochrome P450, 247-322 (Springer US,<br />

Crameri, A., Raillard, S.-A., Bermudez, E. <strong>and</strong> Stemmer, W. P. C. (1998), DNA shuffling<br />

<strong>of</strong> a family <strong>of</strong> genes from diverse species accelerates directed evolution, Nature<br />

391, 288-91.<br />

Črešnar, B. <strong>and</strong> Petrič, Š. Cytochrome P450 enzymes in the <strong>fungal</strong> kingdom, BBA-Protein<br />

Struct M 1814, 29-35.<br />

Cruz-Morató, C.et al. (2012), Biodegradation <strong>of</strong> <strong>pharmaceuticals</strong> <strong>by</strong> fungi <strong>and</strong> metabolites<br />

identification<br />

1-49 (Springer Berlin / Heidelberg,<br />

Dallacosta, C., Monzani, E. <strong>and</strong> Casella, L. (2003), Reactivity study on microperoxidase-8,<br />

J. Biol. Inorg. Chem. 8, 770-76.<br />

127


128<br />

REFERENCES<br />

Damsten, M. C., van Vugt-Lussenburg, B. M. A., Zeldenthuis, T., de Vlieger, J. S. B.,<br />

Comm<strong>and</strong>eur, J. N. M. <strong>and</strong> Vermeulen, N. P. E. (2008), Application <strong>of</strong> drug<br />

metabolising mutants <strong>of</strong> cytochrome P450 BM3 (CYP102A1) as biocatalysts for the<br />

generation <strong>of</strong> reactive metabolites, Chem-Biol Interact 171, 96-107.<br />

Daughton, C. G. <strong>and</strong> Ternes, T. A. (1999), Pharmaceuticals <strong>and</strong> personal care products in<br />

the environment: agents <strong>of</strong> subtle change?, Environ Health Perspectiv 107, 907-38.<br />

Davey, C. A. <strong>and</strong> Fenna, R. E. (1996), 2.3 Å resolution X-ray crystal structure <strong>of</strong> the<br />

bisubstrate analogue inhibitor salicylhydroxamic acid bound to human<br />

myeloperoxidase: a model for a prereaction complex with hydrogen peroxide,<br />

Biochemistry 35, 10967-73.<br />

Davies, M. B. (1984), The copper(II) catalyzed reaction <strong>of</strong> L-ascorbinc acid with<br />

tris(oxalato)cobaltate(III) ion in aqueous solution, Inorg Chim Acta 92, 141-46.<br />

Dawson, J. H. <strong>and</strong> Sono, M. (1987), Cytochrome P-450 <strong>and</strong> chloroperoxidase: thiolateligated<br />

heme enzymes. Spectroscopic determination <strong>of</strong> their active-site structures<br />

<strong>and</strong> mechanistic implications <strong>of</strong> thiolate ligation, Chemical Rev 87, 1255-76.<br />

Degtyarenko, K. (2004), EC numbers for P450 enzymes, website,<br />

Desjardins, A. E. <strong>and</strong> VanEtten, H. D. (1986), Partial purification <strong>of</strong> pisatin demethylase,<br />

a cytochrome P-450 from the pathogenic fungus Nectria haematococca, Arch<br />

Microbiol 144, 84-90.<br />

Deutsche Sammlung von Mikroorganismen und Zellkulturen GmbH, L.-I. D. (2012),<br />

Coprinus radians (Desmazieres) Fries<br />

Ding, X. <strong>and</strong> Kaminsky, L. S. (2003), Human extrahepatic cytochromes P450: function in<br />

xenobiotic metabolism <strong>and</strong> tissue-selective chemical toxicity in the respiratory <strong>and</strong><br />

gastrointestinal tracts, Annu Rev Pharmacol 43, 149-73.<br />

Doddapaneni, H., Chakraborty, R. <strong>and</strong> Yadav, J. S. (2005), Genome-wide structural <strong>and</strong><br />

evolutionary analysis <strong>of</strong> the P450 monooxygenase genes (P450ome) in the white rot<br />

fungus Phanerochaete chrysosporium : Evidence for gene duplications <strong>and</strong><br />

extensive gene clustering, BMC Genomics 6<br />

Dorovska-Taran, V., Posthumus, M. A., Boeren, S., Boersma, M. G., Teunis, C. J.,<br />

Rietjens, I. M. <strong>and</strong> Veeger, C. (1998), Oxygen exchange with water in heme-oxo<br />

intermediates during H 2 O 2 -driven oxygen incorporation in aromatic hydrocarbons<br />

catalyzed <strong>by</strong> microperoxidase-8., Eur. J. Biochem. 253, 659-568.<br />

Dzialoszyński , L. (1955), 4-Nitrocatechol sulfate as a substrate for the assay <strong>of</strong> aryl<br />

sulfatase, Acta. Biochim. Pol. 2, 421-28.<br />

Eiben, S., Kaysser, L., Maurer, S., Kuhnel, K., Urlacher, V. B. <strong>and</strong> Schmid, R. D. (2006),<br />

Preparative use <strong>of</strong> isolated CYP102 monooxygenases-A critical appraisal, J<br />

BIOTECHNOL 124, 662-9.<br />

ElSohly, M. A. <strong>and</strong> Feng, S. (1998), Δ9-THC Metabolites in Meconium: Identification <strong>of</strong><br />

11-OH-Δ9-THC, 8β,11-diOH-Δ9-THC, <strong>and</strong> 11-nor-Δ9-THC-9-COOH as Major<br />

Metabolites <strong>of</strong> Δ9-THC J Anal Toxicol 22, 329-35.<br />

Ferris, J. P., MacDonald, L. H., Patrie, M. A. <strong>and</strong> Martin, M. A. (1976), Aryl hydrocarbon<br />

hydroxylase activity in the fungus Cunninghamella bainieri: Evidence for the<br />

presence <strong>of</strong> cytochrome P-450, Arch Biochem Biophys 175, 443-52.<br />

Feyereisen, R. (1999), Insect P450 enzymes, Annu Rev Entomol 44, 507-33.<br />

Feyereisen, R. (2006), Evolution <strong>of</strong> insect P450, Biochem Soc Trans 34, 1252-5.<br />

Fiedler, T. J., Davey, C. A. <strong>and</strong> Fenna, R. E. (2000), X-ray crystal structure <strong>and</strong><br />

characterization <strong>of</strong> halide-binding sites <strong>of</strong> human myeloperoxidase at 1.8 Ǻ<br />

resolution, J BIOL CHEM 275, 11964-71.<br />

Fishbein, L. <strong>and</strong> Falk, H. L. (1969), Environmental aspects <strong>of</strong> methylenedioxyphenyl <strong>and</strong><br />

related derivatives, Environ Res 2, 297-320.<br />

Frear, D. S. (1995), Wheat microsomal cytochrome P450 monooxygenases:<br />

characterization <strong>and</strong> importance in the metabolic detoxification <strong>and</strong> selectivity <strong>of</strong><br />

wheat herbicides, Drug metab drug interact 12, 329-57.


REFERENCES<br />

Fricke, U., Zawinell, A., Zeidan, R. <strong>and</strong> Steden, M. (2012), Anatomisch-therapeutischchemische-Klassifikation<br />

mit tagesdosen. Amtliche fassung des ATC-Index mit<br />

DDD-Angabe für Deutschl<strong>and</strong> im Jahre 2012, Deutsches Institut für Medizinische<br />

Dokumentation und Information 1-220.<br />

Fulco, A. J. (1991), P450BM-3 <strong>and</strong> <strong>other</strong> inducible bacterial P450 cytochromes:<br />

biochemistry <strong>and</strong> regulation, Annu Rev Pharmacol 31, 177-203.<br />

Garl<strong>and</strong>, W. A., Hsiao, K. C., Pantuck, E. J. <strong>and</strong> Conney, A. H. (1977), Quantitative<br />

determination <strong>of</strong> phenacetin <strong>and</strong> its metabolite acetaminophen <strong>by</strong> GLC-chemical<br />

ionization mass spectrometry, J Pharm Sci 66, 340-4.<br />

Garl<strong>and</strong>, W. A., Nelson, S. D. <strong>and</strong> Sasame, H. A. (1976), Primary <strong>and</strong> secondary<br />

deuterium isotope effects in the O-deethylation <strong>of</strong> phenacetin, Biochem Biophys<br />

Res Co 72, 539-45.<br />

German Association for Pharmaceutical Industry (2011), Pharma-Data 2011, German<br />

Association for Pharmaceutical Industry (BPI e.V.),<br />

Gibson, G. G. <strong>and</strong> Skett, P. (2001), Introduction to drug metabolism, viii, 256 p. (Nelson<br />

Thornes Publishers, Cheltenham, UK.<br />

Gibson, G. G. <strong>and</strong> Skett, P. (2008), Introduction to Drug Metabolism, 256 (Nelson<br />

Thornes, London.<br />

Gibson, M., Soper, C. J., Parfitt, R. T. <strong>and</strong> Sewell, G. J. (1984), Studies on the mechanism<br />

<strong>of</strong> microbial N-demethylation <strong>of</strong> codeine <strong>by</strong> cell-free extracts <strong>of</strong> Cunninghamella<br />

bainieri, Enzyme Microb Tech 6, 471-75.<br />

Gillam, E. M. (2008), Engineering cytochrome P450 enzymes, CHEM RES TOXICOL 21,<br />

220-31.<br />

Gorsky, L. D., Koop, D. R. <strong>and</strong> Coon, M. J. (1984), On the stoichiometry <strong>of</strong> the oxidase<br />

<strong>and</strong> monooxygenase reactions catalyzed <strong>by</strong> liver microsomal cytochrome P-450.<br />

Products <strong>of</strong> oxygen reduction, J BIOL CHEM 259, 6812-17.<br />

Greenberg, L. A. <strong>and</strong> Lester, D. (1946), The metabolic fate <strong>of</strong> acetanilid <strong>and</strong> <strong>other</strong> aniline<br />

derivates, J. Pharmacol. Exp. Ther. 88, 87-98.<br />

Gröbe, G., Ullrich, R., Pecyna, M., Kapturska, D., Friedrich, S., H<strong>of</strong>richter, M. <strong>and</strong><br />

Scheibner, K. (2012), High-yield production <strong>of</strong> aromatic peroxygenase <strong>by</strong> the<br />

agaric fungus Marasmius rotula, AMB Express 1, 1-11.<br />

Groves, J. T. (2006), High-valent iron in chemical <strong>and</strong> biological oxidations, J Inorg<br />

Biochem 100, 434-47.<br />

Groves, J. T. <strong>and</strong> McClusky, G. A. (1978), Aliphatic hydroxylation <strong>by</strong> highly purified liver<br />

microsomal cytochrome P-450. Evidence for a carbon radical intermediate,<br />

Biochem Biophys Res Commun 81, 154-60.<br />

Guengerich, F. P. (1987), Oxidative cleavage <strong>of</strong> carboxylic esters <strong>by</strong> cytochrome P-450, J.<br />

Biol. Chem. 262, 8459-62.<br />

Guengerich, F. P. (2001), Common <strong>and</strong> uncommon cytochrome P450 reactions related to<br />

metabolism <strong>and</strong> chemical toxicity, CHEM RES TOXICOL 14, 611-50.<br />

Guengerich, F. P., Hosea, N. A., Parikh, A., Bell-Parikh, L. C., Johnson, W. W., Gillam, E.<br />

M. J. <strong>and</strong> Shimada, T. (1998), Twenty years <strong>of</strong> biochemistry <strong>of</strong> human P450s, Drug<br />

Metab Dispos 26, 1175-78.<br />

Guengerich, F. P. <strong>and</strong> Montellano, P. R. (2005), Human cytochrome P450 enzymes<br />

Cytochrome P450, 377-530 (Springer US,<br />

Hager, L. P., Morris, D. R., Brown, F. S. <strong>and</strong> Eberwein, H. (1966), Chloroperoxidase. II.<br />

Utilization <strong>of</strong> halogen anions., J BIOL CHEM 241, 1769-77.<br />

Hall, P. F. (1986), Cytochromes P-450 <strong>and</strong> the regulation <strong>of</strong> steroid synthesis, Steroids 48,<br />

131-96.<br />

Hamman, M. A., Thompson, G. A. <strong>and</strong> Hall, S. D. (1997), Regioselective <strong>and</strong><br />

stereoselective metabolism <strong>of</strong> ibupr<strong>of</strong>en <strong>by</strong> human cytochrome P450 2C, Biochem.<br />

Pharmacol. 54, 33-41.<br />

129


130<br />

REFERENCES<br />

Hanano, A., Burcklen, M., Flenet, M., Ivancich, A., Louwagie, M., Garin, J. <strong>and</strong> Blee, E.<br />

(2006), Plant seed peroxygenase is an original heme-oxygenase with an EF-h<strong>and</strong><br />

calcium binding motif, J BIOL CHEM 281, 33140-51.<br />

Hannemann, F., Bichet, A., Ewen, K. M. <strong>and</strong> Bernhardt, R. (2007), Cytochrome P450<br />

systems - biological variations <strong>of</strong> electron transport chains, BIOCHIM BIOPHYS<br />

ACTA 1770, 330-44.<br />

Hansen, L. L. <strong>and</strong> Brøsen, K. (1999), Quantitative determination <strong>of</strong> tolbutamide <strong>and</strong> its<br />

metabolites in human plasma <strong>and</strong> urine <strong>by</strong> high-performance liquid<br />

chromatography <strong>and</strong> UV detection, Therap Drug Monitor 21, 664.<br />

Harrison, J. E. <strong>and</strong> Schultz, J. (1976), Studies on the chlorinating activity <strong>of</strong><br />

myeloperoxidase, J BIOL CHEM 251, 1371-4.<br />

Hass, U., Dünnbier, U. <strong>and</strong> Massmann, G. (2012), Occurrence <strong>of</strong> psychoactive compounds<br />

<strong>and</strong> their metabolites in groundwater downgradient <strong>of</strong> a decommissioned sewage<br />

farm in Berlin (Germany), Environ Sci Pollut R 1-11.<br />

Hayhurst, G. P., Harlow, J., Chowdry, J., Gross, E., Hilton, E., Lennard, M. S., Tucker, G.<br />

T. <strong>and</strong> Ellis, S. W. (2001), Influence <strong>of</strong> phenylalanine-481 substitutions on the<br />

catalytic activity <strong>of</strong> cytochrome P450 2D6., Biochem J 355, 373-79.<br />

Heberer, T. (2002), Occurrence, fate, <strong>and</strong> removal <strong>of</strong> pharmaceutical residues in the<br />

aquatic environment: a review <strong>of</strong> recent research data, Toxicol Lett 131, 5-17.<br />

Hennessy, D. J. (1965), Carbamate insecticides, hydride-transferring ability <strong>of</strong><br />

methylenedioxybenzenes as a basis <strong>of</strong> synergistic activity, J Agricul Food Chem 13,<br />

218-20.<br />

Hern<strong>and</strong>ez, A. <strong>and</strong> Ruiz, M. T. (1998), An EXCEL template for calculation <strong>of</strong> enzyme<br />

kinetic parameters <strong>by</strong> non-linear regression, Bioinformatics 14, 227-28.<br />

Hjelmel<strong>and</strong>, L. M., Aronow, L. <strong>and</strong> Trudell, J. R. (1977), Intramolecular determination <strong>of</strong><br />

substituent effects in hydroxylations catalyzed <strong>by</strong> cytochrome P-450, Mol<br />

Pharmacol 13, 634-9.<br />

H<strong>of</strong>fmann, K.-J., Gyllenhaal, O. <strong>and</strong> Vessman, J. (1987), Analysis <strong>of</strong> α-hydroxy<br />

metabolites <strong>of</strong> metoprolol in human urine after phosgene/trimethylsilyl<br />

derivatization, Biol. Mass Spectrom. 14, 543-48.<br />

H<strong>of</strong>fmann, K. J. (1986), Identification <strong>of</strong> the main urinary metabolites <strong>of</strong> omeprazole after<br />

an oral dose to rats <strong>and</strong> dogs, Drug Metab Dispos 14, 341-48.<br />

H<strong>of</strong>richter, M., Scheibner, K., Schneegaß, I. <strong>and</strong> Fritsche, W. (1998), Enzymatic<br />

combustion <strong>of</strong> aromatic <strong>and</strong> aliphatic compounds <strong>by</strong> manganese peroxidase from<br />

Nematoloma frowardii, Appl Environ Microbiol 64, 399-404.<br />

H<strong>of</strong>richter, M. <strong>and</strong> Ullrich, R. (2006), Heme-thiolate haloperoxidases: versatile<br />

biocatalysts with biotechnological <strong>and</strong> environmental significance, Appl Microbiol<br />

Biot 71, 276-88.<br />

H<strong>of</strong>richter, M. <strong>and</strong> Ullrich, R. (2010), New trends in <strong>fungal</strong> biooxidation. Industrial<br />

Applications, in: Mycota X, Vol. 10, pp. 425-49 (Esser, K., Ed.) Springer Berlin<br />

Heidelberg,<br />

H<strong>of</strong>richter, M., Ullrich, R., Pecyna, M., Liers, C. <strong>and</strong> Lundell, T. (2010), New <strong>and</strong> classic<br />

families <strong>of</strong> secreted <strong>fungal</strong> heme peroxidases, Appl Microbiol Biot 87, 871-97.<br />

Holmes, E.et al. (2007), Detection <strong>of</strong> urinary drug metabolite (xenometabolome)<br />

signatures in molecular epidemiology studies via statistical total correlation (NMR)<br />

spectroscopy, Anal. Chem. 79, 2629-40.<br />

Hrycay, E. G. <strong>and</strong> B<strong>and</strong>iera, S. M. (2012), The monooxygenase, peroxidase, <strong>and</strong><br />

peroxygenase properties <strong>of</strong> cytochrome P450, Arch Biochem Biophys 522, 71-89.<br />

Hu, S. <strong>and</strong> Hager, L. P. (1999), Asymmetric epoxidation <strong>of</strong> functionalized cis-olefins<br />

catalyzed <strong>by</strong> chloroperoxidase, Tetrahedron Lett 40, 1641-44.<br />

Hyl<strong>and</strong>, R., Roe, E. G. H., Jones, B. C. <strong>and</strong> Smith, D. A. (2001), Identification <strong>of</strong> the<br />

cytochrome P450 enzymes involved in the N-demethylation <strong>of</strong> sildenafil, Br J Clin<br />

Pharmacol 51, 239-48.


REFERENCES<br />

Ichinose, H., Wariishi, H. <strong>and</strong> Tanaka, H. (2002), Identification <strong>and</strong> characterization <strong>of</strong><br />

novel cytochrome P450 genes from the white-rot basidiomycete Coriolus<br />

versicolor, Appl Microbiol Biot 58, 97-105.<br />

Iida, T., Sumita, T., Ohta, A. <strong>and</strong> Takagi, M. (2000), The cytochrome P450ALK multigene<br />

family <strong>of</strong> an n-alkane-assimilating yeast, Yarrowia lipolytica: cloning <strong>and</strong><br />

characterization <strong>of</strong> genes coding for new CYP52 family members, Yeast 16, 1077-<br />

87.<br />

Imai, Y., Matsunaga, I., Kusunose, E. <strong>and</strong> Ichihara, K. (2000), Unique heme environment<br />

at the putative distal region <strong>of</strong> hydrogen peroxide-dependent fatty acid -<br />

hydroxylase from Sphingomonas paucimobilis (peroxygenase P450 SPα ), J Biochem<br />

128, 189-94.<br />

Isin, E. M. <strong>and</strong> Guengerich, F. P. (2007), Complex reactions catalyzed <strong>by</strong> cytochrome<br />

P450 enzymes, BBA - Gen Subjects 1770, 314-29.<br />

Jacqz-Aigrain, E., Funck-Brentano, C. <strong>and</strong> Cresteil, T. (1993), CYP2D6- <strong>and</strong> CYP3Adependent<br />

metabolism <strong>of</strong> dextromethorphan in humans, Pharmacogenetics 3, 197-<br />

204.<br />

Jin, Y.et al. (2005), CYP2D6 Genotype, antidepressant use, <strong>and</strong> tamoxifen metabolism<br />

during adjuvant breast cancer treatment, J Natl Cancer Inst 97, 30-39.<br />

Johansson, T., Weidolf, L. <strong>and</strong> Jurva, U. (2007), Mimicry <strong>of</strong> phase I drug metabolism –<br />

novel methods for metabolite characterization <strong>and</strong> synthesis, Rapid Commun Mass<br />

Sp 21, 2323-31.<br />

Joo, H., Lin, Z. <strong>and</strong> Arnold, F. H. (1999), Laboratory evolution <strong>of</strong> peroxide-mediated<br />

cytochrome P450 hydroxylation, Nature 399, 670-73.<br />

Joseph-Horne, T. <strong>and</strong> Hollomon, D. W. (1997), Molecular mechanisms <strong>of</strong> azole resistance<br />

in fungi, FEMS Microbiol Lett 149, 141-49.<br />

Julsing, M. K., Cornelissen, S., Bühler, B. <strong>and</strong> Schmid, A. (2008), Heme-iron oxygenases:<br />

powerful industrial biocatalysts?, Curr Opin Chem Biol 12, 177-86.<br />

Jung, S. T., Lauchli, R. <strong>and</strong> Arnold, F. H. Cytochrome P450: taming a wild type enzyme,<br />

Curr Opin Biotech 22, 809-17.<br />

Jux, U., Baginski, R. M., Arnold, H.-G., Krönke, M. <strong>and</strong> Seng, P. N. (2002), Detection <strong>of</strong><br />

pharmaceutical contaminations <strong>of</strong> river, pond, <strong>and</strong> tap water from Cologne<br />

(Germany) <strong>and</strong> surroundings, Int J Hyg Envir Heal 205, 393-98.<br />

Karlson, U., Dwyer, D. F., Hooper, S. W., Moore, E. R., Timmis, K. N. <strong>and</strong> Eltis, L. D.<br />

(1993), Two independently regulated cytochromes P-450 in a Rhodococcus<br />

rhodochrous strain that degrades 2-ethoxyphenol <strong>and</strong> 4-methoxybenzoate, J<br />

Bacteriol 175, 1467-74.<br />

Kasai, N.et al. Metabolism <strong>of</strong> mono- <strong>and</strong> dichloro-dibenzo-p-dioxins <strong>by</strong> Phanerochaete<br />

chrysosporium cytochromes P450, Appl Microbiol Biot 86, 773-80.<br />

Katagiri, M., Ganguli, B. N. <strong>and</strong> Gunsalus, I. C. (1968), A soluble cytochrome P-450<br />

functional in methylene hydroxylation, J BIOL CHEM 243, 3543-46.<br />

Kellermann, G. H. <strong>and</strong> Luyten-Kellermann, M. (1979), On the regulation <strong>of</strong> antipyrine <strong>and</strong><br />

oxazepam metabolism in man., Res Commun Chem Pathol Pharmacol 23, 287-96.<br />

Kellner, D. G., Hung, S.-C., Weiss, K. E. <strong>and</strong> Sligar, S. G. (2002), Kinetic characterization<br />

<strong>of</strong> compound I formation in the thermostable cytochrome P450 CYP119, J BIOL<br />

CHEM 277, 9641-44.<br />

Kelly, S., Kelly, D., Jackson, C., Warrilow, A., Lamb, D. <strong>and</strong> Ortiz de Montellano, P. R.<br />

(2005), The Diversity <strong>and</strong> importance <strong>of</strong> microbial cytochromes P450, in:<br />

Cytochrome P450: Structure, Mechanism, <strong>and</strong> Biochemistry, Vol. 585-617<br />

(Springer US,<br />

Khan, M. M. T. <strong>and</strong> Martell, A. E. (1967), Metal ion <strong>and</strong> metal chelate catalyzed oxidation<br />

<strong>of</strong> ascorbic acid <strong>by</strong> molecular oxygen. I. Cupric <strong>and</strong> ferric ion catalyzed oxidation,<br />

J Am Chem Soc 89, 4176-85.<br />

131


REFERENCES<br />

Kilpatrick, G. J. <strong>and</strong> Smith, T. W. (2005), Morphine-6-glucuronide: Actions <strong>and</strong><br />

mechanisms, Med Res Rev 25, 521-44.<br />

Kinne, M. (2010), The extracellular peroxygenase <strong>of</strong> the agaric fungus Agrocybe aegerita:<br />

catalytic properties <strong>and</strong> physiological background with particular emphasis on<br />

ether cleavage, Qucosa, <strong>Zittau</strong>.<br />

Kinne, M., Poraj-Kobielska, M., Ar<strong>and</strong>a, E., Ullrich, R., Hammel, K. E., Scheibner, K. <strong>and</strong><br />

H<strong>of</strong>richter, M. (2009a), Regioselective preparation <strong>of</strong> 5-hydroxypropranolol <strong>and</strong><br />

4'-hydroxydicl<strong>of</strong>enac with a <strong>fungal</strong> peroxygenase, Bioorg Med Chem Lett 19,<br />

3085-87.<br />

Kinne, M., Poraj-Kobielska, M., Ralph, S. A., Ullrich, R., H<strong>of</strong>richter, M. <strong>and</strong> Hammel, K.<br />

E. (2009b), Oxidative cleavage <strong>of</strong> diverse ethers <strong>by</strong> an extracellular <strong>fungal</strong><br />

peroxygenase, J BIOL CHEM 284, 29343-49.<br />

Kinne, M., Ullrich, R., Hammel, K. E., Scheibner, K. <strong>and</strong> H<strong>of</strong>richter, M. (2008),<br />

Regioselective preparation <strong>of</strong> (R)-2-(4-hydroxyphenoxy)propionic acid with a<br />

<strong>fungal</strong> peroxygenase, Tetrahedron Lett 49, 5950-53.<br />

Kinne, M., Zeisig, C., Ullrich, R., Kayser, G., Hammel, K. E. <strong>and</strong> H<strong>of</strong>richter, M. (2010),<br />

Stepwise oxygenations <strong>of</strong> toluene <strong>and</strong> 4-nitrotoluene <strong>by</strong> a <strong>fungal</strong> peroxygenase,<br />

Biochem. Biophys. Res. Commun. 397, 18-21.<br />

Kleban<strong>of</strong>f, S. J. (1999), Myeloperoxidase, P Assoc Am Physician 111, 383-89.<br />

Kleban<strong>of</strong>f, S. J. (2005), Myeloperoxidase: friend <strong>and</strong> foe, J Leukocyte Biol 77, 598-625.<br />

Kluge, M., Ullrich, R., Dolge, C., Scheibner, K. <strong>and</strong> H<strong>of</strong>richter, M. (2009), Hydroxylation<br />

<strong>of</strong> naphthalene <strong>by</strong> aromatic peroxygenase from Agrocybe aegerita proceeds via<br />

oxygen transfer from H 2 O 2 <strong>and</strong> intermediary epoxidation, Appl. Microbiol.<br />

Biotechnol. 81, 1071-76.<br />

Kluge, M., Ullrich, R., Scheibner, K. <strong>and</strong> H<strong>of</strong>richter, M. (2007), Spectrophotometric assay<br />

for detection <strong>of</strong> aromatic hydroxylation catalyzed <strong>by</strong> <strong>fungal</strong> haloperoxidase–<br />

peroxygenase, Appl Microbiol Biot 75, 1473-78.<br />

Kluge, M., Ullrich, R., Scheibner, K. <strong>and</strong> H<strong>of</strong>richter, M. (2012), Stereoselective benzylic<br />

hydroxylation <strong>of</strong> alkylbenzenes <strong>and</strong> epoxidation <strong>of</strong> styrene derivatives catalyzed <strong>by</strong><br />

the peroxygenase <strong>of</strong> Agrocybe aegerita, Green Chem 14, 440-46.<br />

Ku, H. Y.et al. (2008), The contributions <strong>of</strong> cytochromes P450 3A4 <strong>and</strong> 3A5 to the<br />

metabolism <strong>of</strong> the phosphodiesterase type 5 inhibitors sildenafil, udenafil, <strong>and</strong><br />

vardenafil, Drug Metab Dispos 36, 986-90.<br />

Lah, L.et al. The versatility <strong>of</strong> the <strong>fungal</strong> cytochrome P450 monooxygenase system is<br />

instrumental in xenobiotic detoxification, Mol Microbiol 81, 1374-89.<br />

Law, M. Y., Charles, S. A. <strong>and</strong> Halliwell, B. (1983), Glutathione <strong>and</strong> ascorbic acid in<br />

spinach (Spinacia oleracea) chloroplasts. The effect <strong>of</strong> hydrogen peroxide <strong>and</strong> <strong>of</strong><br />

Paraquat., Biochem J 210 899–903.<br />

Lee, D.-S.et al. (2003), Substrate recognition <strong>and</strong> molecular mechanism <strong>of</strong> fatty acid<br />

hydroxylation <strong>by</strong> cytochrome P450 from Bacillus subtilis, J BIOL CHEM 278,<br />

9761-67.<br />

Lee, S. H.et al. (2010), Decreased detoxification genes <strong>and</strong> genome size make the human<br />

body louse an efficient model to study xenobiotic metabolism, Insect Mol Biol 19,<br />

599-615.<br />

Leemann, T., Transon, C. <strong>and</strong> Dayer, P. (1993), Cytochrome P450TB (CYP2C): A major<br />

monooxygenase catalyzing dicl<strong>of</strong>enac 4'-hydroxylation in human liver, Life<br />

Sciences 52, 29-34.<br />

Lemberger, L. (1973), Tetrahydrocannabinol metabolism in man, Drug Metab Dispos 1,<br />

461-68.<br />

Lequeu, J., Fauconnier, M.-L., Chammaï, A., Bronner, R. <strong>and</strong> Blée, E. (2003), Formation<br />

<strong>of</strong> plant cuticle: evidence for the occurrence <strong>of</strong> the peroxygenase pathway, Plant J<br />

36, 155-64.<br />

132


REFERENCES<br />

Lewis, D. F. V. (1996), Cytochromes P450: structure, function <strong>and</strong> mechanism., Taylor &<br />

Francis, London.<br />

Lewis, D. F. V. (2001), Guide to Cytochromes P450. Structure <strong>and</strong> Function., Taylor &<br />

Francis, London <strong>and</strong> New York.<br />

Lewis, D. F. V. <strong>and</strong> Wiseman, A. (2005), A selective review <strong>of</strong> bacterial forms <strong>of</strong><br />

cytochrome P450 enzymes, Enzyme Microb Tech 36, 377-84.<br />

Li, A. P., Lu, C., Brent, J. A., Pham, C., Fackett, A., Ruegg, C. E. <strong>and</strong> Silber, P. M. (1999),<br />

Cryopreserved human hepatocytes: characterization <strong>of</strong> drug-metabolizing activities<br />

<strong>and</strong> applications in higher throughput screening assays for hepatotoxicity,<br />

metabolic stability, <strong>and</strong> drug–drug interaction potential, Chem-Biol Interact 121,<br />

17-35.<br />

Li, X., Schuler, M. A. <strong>and</strong> Berenbaum, M. R. (2007), Molecular Mechanisms <strong>of</strong> Metabolic<br />

Resistance to Synthetic <strong>and</strong> Natural Xenobiotics, Annu Rev Entomol 52, 231-53.<br />

Lienert, J., Güdel, K. <strong>and</strong> Escher, B. I. (2007), Screening Method for Ecotoxicological<br />

Hazard Assessment <strong>of</strong> 42 Pharmaceuticals Considering Human Metabolism <strong>and</strong><br />

Excretory Routes, Environ Sci Technol 41, 4471-78.<br />

Lindegardh, N., Davies, G. R., Hien, T. T., Farrar, J., Singhasivanon, P., Day, N. P. J. <strong>and</strong><br />

White, N. J. (2006), Rapid degradation <strong>of</strong> oseltamivir phosphate in clinical samples<br />

<strong>by</strong> plasma esterases, Antimicrob Agents Ch 50, 3197-99.<br />

Liu, G., Gelboin, H. V. <strong>and</strong> Myers, M. J. (1991), Role <strong>of</strong> cytochrome P450 IA2 in<br />

acetanilide 4-hydroxylation as determined with cDNA expression <strong>and</strong> monoclonal<br />

antibodies, Arch Biochem Biophys 284, 400-6.<br />

Lu, A. Y. H. <strong>and</strong> Coon, M. J. (1968), Role <strong>of</strong> Hemoprotein P-450 in Fatty Acid ω-<br />

Hydroxylation in a Soluble Enzyme System from Liver Microsomes J BIOL CHEM<br />

243, 1331-32.<br />

MacKeith, R. A., McCague, R., Olivo, H. F., Roberts, S. M., Taylor, S. J. C. <strong>and</strong> Xiong, H.<br />

(1994), Enzyme-catalysed kinetic resolution <strong>of</strong> 4-endo-hydroxy-2-<br />

oxabicyclo[3.3.0]oct-7-en-3-one <strong>and</strong> employment <strong>of</strong> the pure enantiomers for the<br />

synthesis <strong>of</strong> anti-viral <strong>and</strong> hypocholestemic agents, Bioorg Med Chem 2, 387-94.<br />

Mahato, S. B. <strong>and</strong> Majumdar, I. (1993), Current trends in microbial steroid<br />

biotransformation, Phytochemistry 34, 883-98.<br />

Malonek, S., Rojas, M. C., Hedden, P., Gaskin, P., Hopkins, P. <strong>and</strong> Tudzynski, B. (2005),<br />

Functional Characterization <strong>of</strong> Two Cytochrome P450 Monooxygenase Genes,<br />

P450-1 <strong>and</strong> P450-4, <strong>of</strong> the Gibberellic Acid Gene Cluster in Fusarium proliferatum<br />

(Gibberella fujikuroi MP-D), Appl Environ Microbiol 71, 1462-72.<br />

Maloney, A. P. <strong>and</strong> VanEtten, H. D. (1994), A gene from the <strong>fungal</strong> plant pathogen Nectria<br />

haematococca that encodes the phytoalexin-detoxifying enzyme pisatin demethylase<br />

defines a new cytochrome P450 family, Mol Gen Genet 243, 506-14.<br />

Mansuy, D. (1998), The great diversity <strong>of</strong> reactions catalyzed <strong>by</strong> cytochromes P450, Comp<br />

Biochem Phys C 121, 5-14.<br />

Marco-Urrea, E., Perez-Trujillo, M., Cruz-Morato, C., Caminal, G. <strong>and</strong> Vicent, T. (2010a),<br />

Degradation <strong>of</strong> the drug sodium dicl<strong>of</strong>enac <strong>by</strong> Trametes versicolor pellets <strong>and</strong><br />

identification <strong>of</strong> some intermediates <strong>by</strong> NMR, J Hazard Mater 176, 836-42.<br />

Marco-Urrea, E., Pérez-Trujillo, M., Cruz-Morató, C., Caminal, G. <strong>and</strong> Vicent, T. (2010b),<br />

White-rot fungus-mediated degradation <strong>of</strong> the analgesic ketopr<strong>of</strong>en <strong>and</strong><br />

identification <strong>of</strong> intermediates <strong>by</strong> HPLC–DAD–MS <strong>and</strong> NMR, Chemosphere 78,<br />

474-81.<br />

Marco-Urrea, E., Perez-Trujillo, M., Vicent, T. <strong>and</strong> Caminal, G. (2009), Ability <strong>of</strong> whiterot<br />

fungi to remove selected <strong>pharmaceuticals</strong> <strong>and</strong> identification <strong>of</strong> degradation<br />

products <strong>of</strong> ibupr<strong>of</strong>en <strong>by</strong> Trametes versicolor, Chemosphere 74, 765-72.<br />

Marco-Urrea, E., Radjenovic, J., Caminal, G., Petrovic, M., Vicent, T. <strong>and</strong> Barcelo, D.<br />

(2010c), Oxidation <strong>of</strong> atenolol, propranolol, carbamazepine <strong>and</strong> cl<strong>of</strong>ibric acid <strong>by</strong> a<br />

133


REFERENCES<br />

biological Fenton-like system mediated <strong>by</strong> the white-rot fungus Trametes<br />

versicolor, Water Res 44, 521-32.<br />

Martinez Bueno, M. J., Agüera, A., Gómez, M. J., Hern<strong>and</strong>o, M. D., García-Reyes, J. F.<br />

<strong>and</strong> Fernández-Alba, A. R. (2007), Application <strong>of</strong> Liquid<br />

Chromatography/Quadrupole-Linear Ion Trap Mass Spectrometry <strong>and</strong> Time-<strong>of</strong>-<br />

Flight Mass Spectrometry to the Determination <strong>of</strong> Pharmaceuticals <strong>and</strong> Related<br />

Contaminants in Wastewater, Anal Chem 79, 9372-84.<br />

Martinović, D., Hogarth, W. T., Jones, R. E. <strong>and</strong> Sorensen, P. W. (2007), Environmental<br />

estrogens suppress hormones, behavior, <strong>and</strong> reproductive fitness in male fathead<br />

minnows, Environ Toxicol Chem 26, 271-78.<br />

Masaphy, S., Levanon, D., Henis, Y., Venkateswarlu, K. <strong>and</strong> Kelly, S. L. (1995),<br />

Microsomal <strong>and</strong> cytosolic cytochrome P450 mediated benzo(a)pyrene<br />

hydroxylation in Pleurotus pulmonarius, Biotechnol Lett 17, 969-74.<br />

Maspahy, S., Lamb, D. C. <strong>and</strong> Kelly, S. L. (1999), Purification <strong>and</strong> Characterization <strong>of</strong> a<br />

Benzo[a]pyrene Hydroxylase from Pleurotus pulmonarius, Biochem Biophys Res<br />

Co 266, 326-29.<br />

Masubuchi, Y., Hosokawa, S., Horie, T., Suzuki, T., Ohmori, S., Kitada, M. <strong>and</strong><br />

Narimatsu, S. (1994), Cytochrome P450 isozymes involved in propranolol<br />

metabolism in human liver microsomes. The role <strong>of</strong> CYP2D6 as ring-hydroxylase<br />

<strong>and</strong> CYP1A2 as N-desisopropylase, Drug Metab Dispos 22, 909-15.<br />

Matsuzaki, F. <strong>and</strong> Wariishi, H. (2004), Functional diversity <strong>of</strong> cytochrome P450s <strong>of</strong> the<br />

white-rot fungus Phanerochaete chrysosporium, Biochem Bioph Res Co 324, 387-<br />

93.<br />

Matsuzaki, F. <strong>and</strong> Wariishi, H. (2005), Molecular characterization <strong>of</strong> cytochrome P450<br />

catalyzing hydroxylation <strong>of</strong> benzoates from the white-rot fungus Phanerochaete<br />

chrysosporium, Biochem Bioph Res Co 334, 1184-90.<br />

Maurel, P. (1996), The use <strong>of</strong> adult human hepatocytes in primary culture <strong>and</strong> <strong>other</strong> in<br />

vitro systems to investigate drug metabolism in man, Adv Drug Deliv Rev 22, 105-<br />

32.<br />

McClellan, K. <strong>and</strong> Perry, C. M. (2001), Oseltamivir: A review <strong>of</strong> its use in influenza, Drugs<br />

61, 263-83.<br />

Meunier, B., de Visser, S. P. <strong>and</strong> Shaik, S. (2004), Mechanism <strong>of</strong> Oxidation Reactions<br />

Catalyzed <strong>by</strong> Cytochrome P450 Enzymes, Chem Rev 104, 3947-80.<br />

Meyer, U. A. <strong>and</strong> Zanger, U. M. (1997), Molecular mechanisms <strong>of</strong> genetic polymorphisms<br />

<strong>of</strong> drug metabolism, Ann Rev Pharmacol 37, 269-96.<br />

Miners, J. O. <strong>and</strong> Birkett, D. J. (1996), Use <strong>of</strong> tolbutamide as a substrate probe for human<br />

hepatic cytochrome P450 2C9, Methods Enzymol 272, 139-45.<br />

Miners, J. O., Coulter, S., Tukey, R. H., Veronese, M. E. <strong>and</strong> Birkett, D. J. (1996),<br />

Cytochromes P450, 1A2, <strong>and</strong> 2C9 are responsible for the human hepatic O-<br />

demethylation <strong>of</strong> R- <strong>and</strong> S-naproxen, Biochem Pharmacol 51, 1003-8.<br />

Miyamoto, G., Zahid, N. <strong>and</strong> Uetrecht, J. P. (1997), Oxidation <strong>of</strong> Dicl<strong>of</strong>enac to Reactive<br />

Intermediates <strong>by</strong> Neutrophils, Myeloperoxidase, <strong>and</strong> Hypochlorous Acid, CHEM<br />

RES TOXICOL 10, 414-19.<br />

Mompelat, S., Le Bot, B. <strong>and</strong> Thomas, O. (2009), Occurrence <strong>and</strong> fate <strong>of</strong> pharmaceutical<br />

products <strong>and</strong> <strong>by</strong>-products, from resource to drinking water, Environ Int 35, 803-14.<br />

Muirhead, G. J., Rance, D. J., Walker, D. K. <strong>and</strong> Wastall, P. (2002), Comparative human<br />

pharmacokinetics <strong>and</strong> metabolism <strong>of</strong> single-dose oral <strong>and</strong> intravenous sildenafil,<br />

Brit J Clin Pharmacol 53, 13S-20S.<br />

Munro, A. W., Girvan, H. M. <strong>and</strong> McLean, K. J. (2007), Cytochrome P450 - redox partner<br />

fusion enzymes, BBA - General Subjects 1770, 345-59.<br />

Munro, A. W.et al. (2002), P450 BM3: the very model <strong>of</strong> a modern flavocytochrome,<br />

Trends Biochem Sci 27, 250-57.<br />

134


REFERENCES<br />

Munro, A. W. <strong>and</strong> Lindsay, J. G. (1996), Bacterial cytochromes P-450, Mol Microbiol 20,<br />

1115-25.<br />

Murai, T., Iwabuchi, H. <strong>and</strong> Ikeda, T. (2004), Identification <strong>of</strong> Gemfibrozil Metabolites,<br />

Produced as Positional Isomers in Human Liver Microsomes, <strong>by</strong> On-line Analyses<br />

Using Liquid Chromatography/Mass Spectrometry <strong>and</strong> Liquid<br />

Chromatography/Nuclear Magnetic Resonance Spectroscopy, J Mass Spec Soc Jpn<br />

52, 277-83.<br />

Murray, M. (2000), Mechanisms <strong>of</strong> Inhibitory <strong>and</strong> Regulatory Effects <strong>of</strong><br />

Methylenedioxyphenyl Compounds on Cytochrome P450-Dependent Drug<br />

Oxidation, Curr Drug Metab 1, 67-84.<br />

Nagababu, E. <strong>and</strong> Rifkind, J. M. (2004), Heme Degradation <strong>by</strong> Reactive Oxygen Species<br />

Antioxid Redox Sign 6, 967-78.<br />

Nakayama, N., Takemae, A. <strong>and</strong> Shoun, H. (1996), Cytochrome P450foxy, a Catalytically<br />

Self-Sufficient Fatty Acid Hydroxylase <strong>of</strong> the Fungus Fusarium oxysporum, J<br />

Biochem 119, 435-40.<br />

Narhi, L. O. <strong>and</strong> Fulco, A. J. (1987), Identification <strong>and</strong> characterization <strong>of</strong> two functional<br />

domains in cytochrome P-450BM-3, a catalytically self-sufficient monooxygenase<br />

induced <strong>by</strong> barbiturates in Bacillus megaterium, J BIOL CHEM 262, 6683-90.<br />

Nebert, D. W. <strong>and</strong> Gonzalez, F. J. (1987), P450 Genes: Structure, Evolution, <strong>and</strong><br />

Regulation, Ann Rev Biochem 56, 945-93.<br />

Nelson, D. (2009), The Cytochrome P450 Homepage, Human Genomics 4 59-65.<br />

Nelson, D. R. (2011), Progress in tracing the evolutionary paths <strong>of</strong> cytochrome P450,<br />

BBA - Proteins Proteomics 1814, 14-18.<br />

Nelson, D. R.et al. The P450 superfamily: update on new sequences, gene mapping,<br />

accession numbers, early trivial names <strong>of</strong> enzymes, <strong>and</strong> nomenclature, DNA Cell<br />

Biol 12, 1-51.<br />

Nelson, D. R.et al. (1996), P450 superfamily: update on new sequences, gene mapping,<br />

accession numbers <strong>and</strong> nomenclature, Pharmacogenetics 6, 1-42.<br />

Nelson, D. R. <strong>and</strong> Nebert, D. W. (2011), Cytochrome P450 (CYP) gene superfamily, in:<br />

eLS, Vol. John Wiley & Sons, Ltd,<br />

Nelson, S. D. (1982), Metabolic activation <strong>and</strong> drug toxicity, Journal <strong>of</strong> Medicinal<br />

Chemistry 25, 753-65.<br />

Nelson, S. D. <strong>and</strong> Trager, W. F. (2003), The use <strong>of</strong> deuterium isotope effects to probe the<br />

active site properties, mechanism <strong>of</strong> cytochrome P450-catalyzed reactions, <strong>and</strong><br />

mechanisms <strong>of</strong> metabolically dependent toxicity, Drug Metab. Dispos. 31, 1481-98.<br />

Nielsen, K., Møller, B. <strong>and</strong> Ortiz de Montellano, P. R. (2005), Cytochrome P450s in Plants<br />

Cytochrome P450, 553-83 (Springer US,<br />

Nomenclature Committee <strong>of</strong> the International Union <strong>of</strong> Biochemistry <strong>and</strong> Molecular<br />

Biology (NC-IUBMB) (2012), Enzyme Nomenclature,<br />

http://www.chem.qmul.ac.uk/iubmb/enzyme/<br />

Nottebaum, L. J. <strong>and</strong> McClure, T. D. (2010), Methods, systems, <strong>and</strong> computer program<br />

products for producing theoretical mass spectral fragmentation patterns <strong>of</strong><br />

chemical structures, in: ip.com, Vol. Syngenta Participations AG, USA.<br />

Oaks, J. L.et al. (2004), Dicl<strong>of</strong>enac residues as the cause <strong>of</strong> vulture population decline in<br />

Pakistan, Nature 427, 630-33.<br />

Oatis, J. E., Jr., Russell, M. P., Knapp, D. R. <strong>and</strong> Walle, T. (1981), Ring-hydroxylated<br />

propranolol: synthesis <strong>and</strong> beta-receptor antagonist <strong>and</strong> vasodilating activities <strong>of</strong><br />

the seven isomers, J Med Chem 24, 309-14.<br />

Ogliaro, F., de Visser, S. P., Cohen, S., Sharma, P. K. <strong>and</strong> Shaik, S. (2002), Searching for<br />

the second oxidant in the catalytic cycle <strong>of</strong> cytochrome P450: a theoretical<br />

investigation <strong>of</strong> the iron(III)-hydroperoxo species <strong>and</strong> its epoxidation pathways, J<br />

Am Chem Soc 124, 2806-17.<br />

135


136<br />

REFERENCES<br />

Ohkuma, M., Zimmer, T., Iida, T., Schunck, W.-H., Ohta, A. <strong>and</strong> Takagi, M. (1998),<br />

Isozyme Function <strong>of</strong> n-Alkane-inducible Cytochromes P450 in C<strong>and</strong>ida maltosa<br />

Revealed <strong>by</strong> Sequential Gene Disruption, J BIOL CHEM 273, 3948-53.<br />

Omura, T. (2005), Heme-thiolate proteins, Biochem. Biophys. Res. Commun. 338, 404-09.<br />

Omura, T., S<strong>and</strong>ers, E., Estabrook, R. W., Cooper, D. Y. <strong>and</strong> Rosenthal, O. (1966),<br />

Isolation from adrenal cortex <strong>of</strong> a nonheme iron protein <strong>and</strong> a flavoprotein<br />

functional as a reduced triphosphopyridine nucleotide-cytochrome P-450<br />

reductase, Arch Biochem Biophys 117, 660-73.<br />

Omura, T. <strong>and</strong> Sato, R. (1964), The Carbon Monoxide-binding Pigment <strong>of</strong> Liver<br />

Microsomes, J BIOL CHEM 239, 2370-78.<br />

Ophardt, E. C. (2003), Amino acids, in: Virtual ChemBook, Vol. Elmhurst College.<br />

Orl<strong>and</strong>o, R., Piccoli, P., De Martin, S., Padrini, R., Floreani, M. <strong>and</strong> Palatini, P. (2004),<br />

Cytochrome P450 1A2 is a major determinant <strong>of</strong> lidocaine metabolism in vivo:<br />

effects <strong>of</strong> liver function, Clin Pharmacol Ther 75, 80-88.<br />

Ortiz de Montellano, P. (2005), Cytochrome P450 - Structure, Mechanism <strong>and</strong><br />

Biochemistry, Ortiz de Montellano, P., Ed.) Kluwer Academic/Plenum Publishers,<br />

New York.<br />

Ortiz de Montellano, P., Voss, J. <strong>and</strong> Ortiz de Montellano, P. R. (2005), Substrate<br />

Oxidation <strong>by</strong> Cytochrome P450 Enzymes<br />

Cytochrome P450, 183-245 (Springer US,<br />

Ortiz de Montellano, P. R. <strong>and</strong> de Voss, J. J. (2005), Substrate oxidation <strong>by</strong> cytochrome<br />

P450 enzymes, in: Cytochrome P450 - Structure, Mechanism <strong>and</strong> Biochemistry,<br />

Vol. 183-245 (Ortiz De Montellano, P. R., Ed.) Kluwer Academic/Plenum<br />

Publishers, New York.<br />

Osman, A. M., Koerts, J., Boersma, M. G., Boeren, S., Veeger, C. <strong>and</strong> Rietjens, I. M.<br />

(1996), Microperoxidase/H2O2-catalyzed aromatic hydroxylation proceeds <strong>by</strong> a<br />

cytochrome-P-450-type oxygen-transfer reaction mechanism, Eur J Biochem 240,<br />

232-8.<br />

Otey, C. R., B<strong>and</strong>ara, G., Lalonde, J., Takahashi, K. <strong>and</strong> Arnold, F. H. (2006), Preparation<br />

<strong>of</strong> human metabolites <strong>of</strong> propranolol using laboratory-evolved bacterial<br />

cytochromes P450, Biotechnol Bioeng 93, 494-9.<br />

Paice, M. G., Reid, I. D., Bourbonnais, R., Archibald, F. S. <strong>and</strong> Jurasek, L. (1993),<br />

Manganese peroxidase, produced <strong>by</strong> Trametes versicolor during pulp bleaching,<br />

demethylates <strong>and</strong> delignifies kraft pulp, Appl Environ Microbiol 59, 260-65.<br />

Parliment, E., The European Parliment <strong>and</strong> Council, E., The Council Of The European<br />

Union (2004), Directive 2004/27/EC <strong>of</strong> the European Parliament <strong>and</strong> <strong>of</strong> the<br />

Council <strong>of</strong> 31 March 2004 amending directive 2001/83/EC on the community code<br />

relating to medicinal products for human use, 136, pp. 0034 - 57 (Official J,<br />

Pearce, R. E., Vakkalagadda, G. R. <strong>and</strong> Leeder, J. S. (2002), Pathways <strong>of</strong> carbamazepine<br />

bioactivation in vitro I. Characterization <strong>of</strong> human cytochromes P450 responsible<br />

for the formation <strong>of</strong> 2- <strong>and</strong> 3-hydroxylated metabolites, Drug Metab Dispos 30,<br />

1170-9.<br />

Pecyna, M. (2012), Gen sequences <strong>of</strong> <strong>fungal</strong> peroxygenases, oral communication,<br />

International Graduate School <strong>Zittau</strong>, oral communication,<br />

Pecyna, M. J., Ullrich, R., Bittner, B., Clemens, A., Scheibner, K., Schubert, R. <strong>and</strong><br />

H<strong>of</strong>richter, M. (2009), Molecular characterization <strong>of</strong> aromatic peroxygenase from<br />

Agrocybe aegerita, Appl Microbiol Biotechnol 84, 885-97.<br />

Perry, T. L., Culling, C. F. A., Berry, K. <strong>and</strong> Hansen, S. (1964), 7-<br />

Hydroxychlorpromazine: Potential Toxic Drug Metabolite in Psychiatric Patients,<br />

Science 146, 81-83.<br />

Peter, S., Kinne, M., Wang, X., Ullrich, R., Kayser, G., Groves, J. T. <strong>and</strong> H<strong>of</strong>richter, M.<br />

(2011), Selective hydroxylation <strong>of</strong> alkanes <strong>by</strong> an extracellular <strong>fungal</strong> peroxygenase,<br />

FEBS J 278, 3667-75.


REFERENCES<br />

Peters, F. T., Dragan, C. A., Wilde, D. R., Meyer, M. R., Zapp, J., Bureik, M. <strong>and</strong> Maurer,<br />

H. H. (2007), Biotechnological synthesis <strong>of</strong> drug metabolites using human<br />

cytochrome P450 2D6 heterologously expressed in fission yeast exemplified for the<br />

designer drug metabolite 4'-hydroxymethyl-alpha-pyrrolidinobutyrophenone,<br />

Biochem Pharmacol 74, 511-20.<br />

Petrović, M., Gonzalez, S. <strong>and</strong> Barceló, D. (2003), Analysis <strong>and</strong> removal <strong>of</strong> emerging<br />

contaminants in wastewater <strong>and</strong> drinking water, TrAC-Trend Anal Chem 22, 685-<br />

96.<br />

Philpot, R. M. <strong>and</strong> Hodgson, E. (1972), The production <strong>and</strong> modification <strong>of</strong> cytochrome P-<br />

450 difference spectra <strong>by</strong> in vivo administration <strong>of</strong> methylenedioxyphenyl<br />

compounds, Chem-Biol Interact 4, 185-94.<br />

Piontek, K., Ullrich, R., Liers, C., Diederichs, K., Plattner, D. A. <strong>and</strong> H<strong>of</strong>richter, M.<br />

(2010), Crystallization <strong>of</strong> a 45 kDa peroxygenase/peroxidase from the mushroom<br />

Agrocybe aegerita <strong>and</strong> structure determination <strong>by</strong> SAD utilizing only the haem<br />

iron, Acta Crystallogr Sect F Struct Biol Cryst Commun 66, 693-8.<br />

Piontek, K.et al. (2012), Crystal structures <strong>of</strong> aromatic peroxygenase: substrate-binding<br />

mode in a new class <strong>of</strong> heme redox enzymes,<br />

Poraj-Kobielska, M., Kinne, M., Ullrich, R., Scheibner, K. <strong>and</strong> H<strong>of</strong>richter, M. (2011a), A<br />

spectrophotometric assay for the detection <strong>of</strong> <strong>fungal</strong> peroxygenases, Anal Bioch<br />

421, 327-29.<br />

Poraj-Kobielska, M., Kinne, M., Ullrich, R., Scheibner, K., Kayser, G., Hammel, K. E. <strong>and</strong><br />

H<strong>of</strong>richter, M. (2011b), Preparation <strong>of</strong> human drug metabolites using <strong>fungal</strong><br />

peroxygenases, Biochem Pharmacol 82, 789-96.<br />

Prieto, T., Marcon, R. O., Prado, F. M., Caires, A. C., Di Mascio, P., Brochsztain, S.,<br />

Nascimento, O. R. <strong>and</strong> Nantes, I. L. (2006), Reaction route control <strong>by</strong><br />

microperoxidase-9/CTAB micelle ratios, Phys. Chem. Chem. Phys. 8, 1963-73.<br />

Quertemont, E. (2004), Genetic polymorphism in ethanol metabolism: acetaldehyde<br />

contribution to alcohol abuse <strong>and</strong> alcoholism, Mol Psychiatry 9, 570-81.<br />

Quinn, B., Gagne, F. <strong>and</strong> Blaise, C. (2008), An investigation into the acute <strong>and</strong> chronic<br />

toxicity <strong>of</strong> eleven <strong>pharmaceuticals</strong> (<strong>and</strong> their solvents) found in wastewater effluent<br />

on the cnidarian, Hydra attenuata, Sci Total Environ 389, 306-14.<br />

Rittle, J. <strong>and</strong> Green, M. T. (2010), Cytochrome P450 compound I: capture,<br />

characterization, <strong>and</strong> C-H bond activation kinetics, Science 330, 933-7.<br />

Rodarte-Morales, A., Feijoo, G., Moreira, M. <strong>and</strong> Lema, J. (2012), Biotransformation <strong>of</strong><br />

three pharmaceutical active compounds <strong>by</strong> the fungus Phanerochaete<br />

chrysosporium in a fed batch stirred reactor under air <strong>and</strong> oxygen supply,<br />

Biodegradation 23, 145-56.<br />

Roy, A., Kucukural, A. <strong>and</strong> Zhang, Y. (2010), I-TASSER: a unified platform for automated<br />

protein structure <strong>and</strong> function prediction, Nat Protocol 5, 725-38.<br />

Ruiz-Garcia, A., Bermejo, M., Moss, A. <strong>and</strong> Casabo, V. G. (2008), Pharmacokinetics in<br />

drug discovery, J Pharm Sci 97, 654-90.<br />

Sachverständigen Rat für Umweltfragen (2007), Arzneimittel in der Umwelt, 1-95.<br />

Saikia, S., Parker, E. J., Koulman, A. <strong>and</strong> Scott, B. (2007), Defining Paxilline Biosynthesis<br />

in Penicillium paxilli, J BIOL CHEM 282, 16829-37.<br />

Sawayama, A. M., Chen, M. M., Kulanthaivel, P., Kuo, M. S., Hemmerle, H. <strong>and</strong> Arnold,<br />

F. H. (2009), A panel <strong>of</strong> cytochrome P450 BM3 variants to produce drug<br />

metabolites <strong>and</strong> diversify lead compounds, Chem Eur J 15, 11723-9.<br />

Scheibner, K. <strong>and</strong> H<strong>of</strong>richter, M. (1998), <strong>Conversion</strong> <strong>of</strong> aminonitrotoluenes <strong>by</strong> <strong>fungal</strong><br />

manganese peroxidase, Journal <strong>of</strong> Basic Microbiology 38, 51-59.<br />

Scheller, U., Zimmer, T., Kärgel, E. <strong>and</strong> Schunck, W.-H. (1996), Characterization <strong>of</strong> then-<br />

Alkane <strong>and</strong> Fatty Acid Hydroxylating Cytochrome P450 Forms 52A3 <strong>and</strong> 52A4,<br />

Arch Biochem Biophys 328, 245-54.<br />

137


138<br />

REFERENCES<br />

Schenkman, J. B., Cinti, D. L., Orrenius, S., Moldeus, P. <strong>and</strong> Kraschnitz, R. (1972), The<br />

nature <strong>of</strong> the reverse type I (modified type II) spectral change in liver microsomes,<br />

Biochemistry 11, 4243-51.<br />

Schenkman, J. B., Sligar, S. G. <strong>and</strong> Cinti, D. L. (1981), Substrate interaction with<br />

cytochrome P-450, Pharmacol Therapeut 12, 43-71.<br />

Schmidt, H. W. H., Haemmerli, S. D., Schoemaker, H. E. <strong>and</strong> Leisola, M. S. A. (1989),<br />

Oxidative degradation <strong>of</strong> 3,4-dimethoxybenzyl alcohol <strong>and</strong> its methyl ether <strong>by</strong> the<br />

lignin peroxidase <strong>of</strong> Phanerochaete chrysosporium, Biochemistry 28, 1776-83.<br />

Schroer, K., Kittelmann, M. <strong>and</strong> Lütz, S. (2010), Recombinant human cytochrome P450<br />

monooxygenases for drug metabolite synthesis, Biotechnol Bioeng 106, 699-706.<br />

Schuler, M. A. (1996a), Plant Cytochrome P450 Monooxygenases, CRC Cr Rev Plant Sci<br />

15, 235-84.<br />

Schuler, M. A. (1996b), The Role <strong>of</strong> Cytochrome P450 Monooxygenases in Plant-Insect<br />

Interactions, Plant Physiol 112, 1411-19.<br />

Schwarz, H. J. (1979), Pharmacokinetics <strong>and</strong> metabolism <strong>of</strong> temazepam in man <strong>and</strong><br />

several animal species, Br J Clin Pharmacol 8, 23S–29S.<br />

Scott, J. G. (2008), Insect cytochrome P450s: Thinking beyond detoxification, Recent<br />

Advances in Insect Physiology, Toxicology <strong>and</strong> Molecular Biology ISBN: 978-81-<br />

308-0242-8, 117-24.<br />

Segel, H. I. (1994), Enzyme Kinetics: Behavior <strong>and</strong> Analysis <strong>of</strong> Rapid Equilibrium <strong>and</strong><br />

Steady-State Enzyme Systems, John Wiley Sons, Inc., New York<br />

Segre, E. J. (1975), Naproxen metabolism in man, J Clin Pharmacol 15, 316-23.<br />

Shoji, O., Fujishiro, T., Nakajima, H., Kim, M., Nagano, S., Shiro, Y. <strong>and</strong> Watanabe, Y.<br />

(2007), Hydrogen peroxide dependent monooxygenations <strong>by</strong> tricking the substrate<br />

recognition <strong>of</strong> cytochrome P450BSβ, Angew Chem Int Ed 46, 3656-59.<br />

Shoji, O., Wiese, C., Fujishiro, T., Shirataki, C., Wünsch, B. <strong>and</strong> Watanabe, Y. (2010),<br />

Aromatic C–H bond hydroxylation <strong>by</strong> P450 peroxygenases: a facile colorimetric<br />

assay for monooxygenation activities <strong>of</strong> enzymes based on Russig’s blue formation,<br />

J Biol Inorg Chem 15, 1109-15.<br />

Shoun, H., Fushinobu, S., Jiang, L., Kim, S.-W. <strong>and</strong> Wakagi, T. Fungal denitrification <strong>and</strong><br />

nitric oxide reductase cytochrome P450nor, Philos T Roy Soc B 367, 1186-94.<br />

Shoun, H. <strong>and</strong> Tanimoto, T. (1991), Denitrification <strong>by</strong> the fungus Fusarium oxysporum<br />

<strong>and</strong> involvement <strong>of</strong> cytochrome P-450 in the respiratory nitrite reduction, J BIOL<br />

CHEM 266, 11078-82.<br />

Simonneaux, G., Le Maux, P. <strong>and</strong> Simonneaux, G. (2006), Carbene Complexes <strong>of</strong> Heme<br />

Proteins <strong>and</strong> Iron Porphyrin Models<br />

Bioorganometallic Chemistry, 17, pp. 83-122 (Springer Berlin / Heidelberg,<br />

Skordi, E., Wilson, I. D., Lindon, J. C. <strong>and</strong> Nicholson, J. K. (2005), Kinetic studies on the<br />

intramolecular acyl migration <strong>of</strong> β-1-O-acyl glucuronides: Application to the<br />

glucuronides <strong>of</strong> (R)- <strong>and</strong> (S)-ketopr<strong>of</strong>en, (R)- <strong>and</strong> (S)-hydroxy-ketopr<strong>of</strong>en<br />

metabolites, <strong>and</strong> tolmetin <strong>by</strong> 1H-NMR spectroscopy, Xenobiotica 35, 715-25.<br />

Sligar, S. G. <strong>and</strong> Gunsalus, I. C. (1976), A thermodynamic model <strong>of</strong> regulation: modulation<br />

<strong>of</strong> redox equilibria in camphor monoxygenase, P Natl Acad Sci USA 73, 1078-82.<br />

Sono, M., Roach, M. P., Coulter, E. D. <strong>and</strong> Dawson, J. H. (1996), Heme-containing<br />

oxygenases, Chem Rev 96, 2841-88.<br />

Souček, P. <strong>and</strong> Gut, I. (1992), Cytochromes P-450 in rats: structures, functions, properties<br />

<strong>and</strong> relevant human forms, Xenobiotica 22, 83-103.<br />

Stamets, P. (1993), Growing gourmet <strong>and</strong> medicinal mushrooms: a companion guide to<br />

The mushroom cultivator, Ten Speed Press (Berkeley, CA) Berleöey.<br />

Stemmer, W. P. C. (1994), Rapid evolution <strong>of</strong> a protein in vitro <strong>by</strong> DNA shuffling, Nature<br />

370, 389-91.<br />

Strode, C.et al. (2008), Genomic analysis <strong>of</strong> detoxification genes in the mosquito Aedes<br />

aegypti, Insect Biochem Molec 38, 113-23.


REFERENCES<br />

Stundl, U. M., Patzak, D. <strong>and</strong> Schauer, F. (2000), Purification <strong>of</strong> a soluble cytochrome<br />

P450 from Trichosporon montevideense, J Basic Microbiol 40, 289-92.<br />

Subramanian, V., Doddapaneni, H., Syed, K. <strong>and</strong> Yadav, J. (2010), P450 Redox Enzymes<br />

in the White Rot Fungus Phanerochaete chrysosporium: Gene Transcription,<br />

Heterologous Expression, <strong>and</strong> Activity Analysis on the Purified Proteins, Curr<br />

Microbiol 61, 306-14.<br />

Sugano, Y., Muramatsu, R., Ichiyanagi, A., Sato, T. <strong>and</strong> Shoda, M. (2007), DyP, a unique<br />

dye-decolorizing peroxidase, represents a novel heme peroxidase family: ASP171<br />

replaces the distal histidine <strong>of</strong> classical peroxidases, J BIOL CHEM 282, 36652-8.<br />

Sundaramoorthy, M., Terner, J. <strong>and</strong> Poulos, T. L. (1995), The crystal structure <strong>of</strong><br />

chloroperoxidase: a heme peroxidase--cytochrome P450 functional hybrid,<br />

Structure 3, 1367-77.<br />

Sundaramoorthy, M. T., J. Poulos, T.L. (1995), The crystal structure <strong>of</strong> chloroperoxidase:<br />

a heme peroxidase--cytochrome P450 functional hybrid., Structure 3, 1367-77.<br />

Suzuki, K., Sanga, K.-i., Chikaoka, Y. <strong>and</strong> Itagaki, E. (1993), Purification <strong>and</strong> properties<br />

<strong>of</strong> cytochrome P-450 (P-450lun) catalyzing steroid 11β-hydroxylation in<br />

Curvularia lunata, BBA - Protein Struct M 1203, 215-23.<br />

Syed, K., Porollo, A., Lam, Y. W. <strong>and</strong> Yadav, J. S. (2011), A Fungal P450 (CYP5136A3)<br />

Capable <strong>of</strong> Oxidizing Polycyclic Aromatic Hydrocarbons <strong>and</strong> Endocrine<br />

Disrupting Alkylphenols: Role <strong>of</strong> Trp 129 <strong>and</strong> Leu 324 , PLoS ONE 6, e28286.<br />

Syed, K. <strong>and</strong> Yadav, J. S. (2012), P450 monooxygenases (P450ome) <strong>of</strong> the model white rot<br />

fungus Phanerochaete chrysosporium, Crc Cr Rev Microbiol 38, 339-63.<br />

Sztal, T., Chung, H., Berger, S., Currie, P. D., Batterham, P. <strong>and</strong> Daborn, P. J. (2012), A<br />

Cytochrome P450 Conserved in Insects Is Involved in Cuticle Formation, PLoS<br />

ONE 7, e36544.<br />

Tam, Y. K., Ke, J., Coutts, R. T., Wyse, D. G. <strong>and</strong> Gray, M. R. (1990), Quantification <strong>of</strong><br />

Three Lidocaine Metabolites <strong>and</strong> Their Conjugates, Pharm Res 7, 504-07.<br />

Tassaneeyakul, W., Veronese, M. E., Birkett, D. J., Gonzalez, F. J. <strong>and</strong> Miners, J. O.<br />

(1993a), Validation <strong>of</strong> 4-nitrophenol as an in vitro substrate probe for human liver<br />

CYP2E1 using cDNA expression <strong>and</strong> microsomal kinetic techniques, Biochem<br />

Pharmacol 46, 1975-81.<br />

Tassaneeyakul, W., Veronese, M. E., Birkett, D. J. <strong>and</strong> Miners, J. O. (1993b), Highperformance<br />

liquid chromatographic assay for 4-nitrophenol hydroxylation, a<br />

putative cytochrome P-4502E1 activity, in human liver microsomes, J. Chromatogr.<br />

616, 73-8.<br />

Ternes, T. A., Joss, A. <strong>and</strong> Siegrist, H. (2004), Peer Reviewed: Scrutinizing<br />

Pharmaceuticals <strong>and</strong> Personal Care Products in Wastewater Treatment, Environ<br />

Sci Technol 38, 392A-99A.<br />

Thomas, J. L., Duax, W. L., Addlagatta, A., Kacsoh, B., Br<strong>and</strong>t, S. E. <strong>and</strong> Norris, W. B.<br />

(2004), Structure/function aspects <strong>of</strong> human 3beta-hydroxysteroid dehydrogenase,<br />

Mol cell endocrinol 215, 73-82.<br />

Tien, M. <strong>and</strong> Kirk, T. K. (1984), Lignin-degrading enzyme from Phanerochaete<br />

chrysosporium: Purification, characterization, <strong>and</strong> catalytic properties <strong>of</strong> a unique<br />

H 2 O 2 -requiring oxygenase, P Natl Acad Sci USA 81, 2280-84.<br />

Tien, M., Kirk, T. K., Bull, C. <strong>and</strong> Fee, J. A. (1986), Steady-state <strong>and</strong> transient-state<br />

kinetic studies on the oxidation <strong>of</strong> 3,4-dimethoxybenzyl alcohol catalyzed <strong>by</strong> the<br />

ligninase <strong>of</strong> Phanerocheate chrysosporium Burds, J. Biol. Chem. 261, 1687-93.<br />

Topham, J. C. (1970), Relationship between difference spectra <strong>and</strong> metabolism:<br />

Barbiturates, drug interaction <strong>and</strong> species difference, Biochem Pharmacol 19,<br />

1695-701.<br />

Tucker, G. T.et al. (1994), The demethylenation <strong>of</strong> methylenedioxymethamphetamine<br />

(“ecstasy”) <strong>by</strong> debrisoquine hydroxylase (CYP2D6), Biochem Pharmacol 47, 1151-<br />

56.<br />

139


140<br />

REFERENCES<br />

Tudzynski, B., Hedden, P., Carrera, E. <strong>and</strong> Gaskin, P. (2001), The P450-4 Gene <strong>of</strong><br />

Gibberella fujikuroi Encodes ent-Kaurene Oxidase in the Gibberellin Biosynthesis<br />

Pathway, Appl Environ Microbiol 67, 3514-22.<br />

Tuynman, A., Spelberg, J. L., Kooter, I. M., Schoemaker, H. E. <strong>and</strong> Wever, R. (2000),<br />

Enantioselective Epoxidation <strong>and</strong> Carbon - Carbon Bond Cleavage Catalyzed <strong>by</strong><br />

Coprinus cinereus Peroxidase <strong>and</strong> Myeloperoxidase, J BIOL CHEM 275, 3025-30.<br />

Udenfriend, S., Clark, C. T., Axelrod, J. <strong>and</strong> Brodie, B. B. (1954), Ascorbic Acid in<br />

Aromatic Hydroxylation, J BIOL CHEM 208, 731-40.<br />

Uljé, K. (2001), Coprinus radians Snowarski, M., Ed.)<br />

Ullrich, R., Dolge, C., Kluge, M. <strong>and</strong> H<strong>of</strong>richter, M. (2008), Pyridine as novel substrate<br />

for regioselective oxygenation with aromatic peroxygenase from Agrocybe<br />

aegerita, FEBS Lett. 582, 4100-06.<br />

Ullrich, R. <strong>and</strong> H<strong>of</strong>richter, M. (2005), The haloperoxidase <strong>of</strong> the agaric fungus Agrocybe<br />

aegerita hydroxylates toluene <strong>and</strong> naphthalene, FEBS Lett. 579, 6247-50.<br />

Ullrich, R. <strong>and</strong> H<strong>of</strong>richter, M. (2007), Enzymatic hydroxylation <strong>of</strong> aromatic compounds,<br />

Cell Mol Life Sci 64, 271-93.<br />

Ullrich, R., Liers, C., Schimpke, S. <strong>and</strong> H<strong>of</strong>richter, M. (2009), Purification <strong>of</strong><br />

homogeneous forms <strong>of</strong> <strong>fungal</strong> peroxygenase, Biotech J 4, 1619-26.<br />

Ullrich, R., Nüske, J., Scheibner, K., Spantzel, J. <strong>and</strong> H<strong>of</strong>richter, M. (2004), Novel<br />

haloperoxidase from the agaric basidiomycete Agrocybe aegerita oxidizes aryl<br />

alcohols <strong>and</strong> aldehydes, Appl Environ Microbiol 70, 4575-81.<br />

Ullrich, R., Pecyna, M., Kluge, M., Kinne, M., Liers, C., H<strong>of</strong>richter, M. (2008), Some<br />

special reactions <strong>of</strong> Agrocybe aegerita peroxygenase (AaP). in: 8th International<br />

Peroxidase Symposium, Vol. 48 (Tampere, Finl<strong>and</strong>.<br />

Upthagrove, A. L. <strong>and</strong> Nelson, W. L. (2001), Importance <strong>of</strong> amine pKa <strong>and</strong> distribution<br />

coefficient in the metabolism <strong>of</strong> fluorinated propranolol derivatives. Preparation,<br />

identification <strong>of</strong> metabolite regioisomers, <strong>and</strong> metabolism <strong>by</strong> CYP2D6, Drug Metab<br />

Dispos 29, 1377-88.<br />

Urlacher, V. B. <strong>and</strong> Girhard, M. Cytochrome P450 monooxygenases: an update on<br />

perspectives for synthetic application, Trends Biotechnol 30, 26-36.<br />

Urlacher, V. B., Lutz-Wahl, S. <strong>and</strong> Schmid, R. D. (2004), Microbial P450 enzymes in<br />

biotechnology., Appl Microb Biotechnol 64, 317-25.<br />

Vail, R. B., Homann, M. J., Hanna, I. <strong>and</strong> Zaks, A. (2005), Preparative synthesis <strong>of</strong> drug<br />

metabolites using human cytochrome P450s 3A4, 2C9 <strong>and</strong> 1A2 with NADPH-P450<br />

reductase expressed in &lt;i&gt;Escherichia coli&lt;/i&gt, J Ind Microbiol Biot<br />

32, 67-74.<br />

van den Brink, H. M., van Gorcom, R. F. M., van den Hondel, C. A. M. J. J. <strong>and</strong> Punt, P. J.<br />

(1998), Cytochrome P450 Enzyme Systems in Fungi, Fungal Genet Biol 23, 1-17.<br />

Veeger, C. (2002), Does P450-type catalysis proceed through a peroxo-iron intermediate?<br />

A review <strong>of</strong> studies with microperoxidase, J. Inorg. Biochem. 91, 35-45.<br />

Venkatakrishnan, K., Von Moltke, L. L. <strong>and</strong> Greenblatt, D. J. (2001), Human drug<br />

metabolism <strong>and</strong> the cytochromes P450: application <strong>and</strong> relevance <strong>of</strong> in vitro<br />

models, J Clin Pharmacol 41, 1149-79.<br />

Waldon, D. J., Teffera, Y., Colletti, A. E., Liu, J., Zurcher, D., Copel<strong>and</strong>, K. W. <strong>and</strong> Zhao,<br />

Z. (2010), Identification <strong>of</strong> quinone imine containing glutathione conjugates <strong>of</strong><br />

dicl<strong>of</strong>enac in rat bile, Chem ResToxicol 23, 1947-53.<br />

Walker, D. K. (1999), Pharmacokinetics <strong>and</strong> metabolism <strong>of</strong> sildenafil in mouse, rat, rabbit,<br />

dog <strong>and</strong> man, Xenobiotica 29, 297-310.<br />

Walsky, R. L. <strong>and</strong> Obach, R. S. (2004), Validated Assays for Human Cytochrome P450<br />

activities, Drug Metab Dispos 32, 647-60.<br />

Wang, X., Peter, S., Kinne, M., H<strong>of</strong>richter, M. <strong>and</strong> Groves, J. T. (2012), Detection <strong>and</strong><br />

Kinetic Characterization <strong>of</strong> a Highly Reactive Heme - Thiolate Peroxygenase<br />

Compound I, J Am Chem Soc 134, 12897-900.


REFERENCES<br />

Warman, A. J.et al. (2005), Flavocytochrome P450 BM3: an update on structure <strong>and</strong><br />

mechanism <strong>of</strong> a biotechnologically important enzyme, Biochem Soc Trans 33, 747-<br />

53.<br />

Warrington, J. S., Shader, R. I., von Moltke, L. L. <strong>and</strong> Greenblatt, D. J. (2000), In vitro<br />

biotransformation <strong>of</strong> sildenafil (Viagra): identification <strong>of</strong> human cytochromes <strong>and</strong><br />

potential drug interactions, Drug Metab Dispos 28, 392-97.<br />

Watanabe, K., Yamaori, S., Funahashi, T., Kimura, T. <strong>and</strong> Yamamoto, I. (2007),<br />

Cytochrome P450 enzymes involved in the metabolism <strong>of</strong> tetrahydrocannabinols<br />

<strong>and</strong> cannabinol <strong>by</strong> human hepatic microsomes, Life Sci 80, 1415-19.<br />

Weis, R., Winkler, M., Schittmayer, M., Kambourakis, S., Vink, M., Rozzell, J. D. <strong>and</strong><br />

Glieder, A. (2009), A Diversified Library <strong>of</strong> Bacterial <strong>and</strong> Fungal Bifunctional<br />

Cytochrome P450 Enzymes for Drug Metabolite Synthesis, Adv Synth Catal 351,<br />

2140-46.<br />

Werck-Reichhart, D. <strong>and</strong> Feyereisen, R. (2000), Cytochromes P450: a success story,<br />

Genome Biol 1, reviews3003.1 - reviews03.9.<br />

Wester, M. R., Johnson, E. F., Marques-Soares, C., Dijols, S., Dansette, P. M., Mansuy, D.<br />

<strong>and</strong> Stout, C. D. (2003), Structure <strong>of</strong> Mammalian Cytochrome P450 2C5<br />

Complexed with Dicl<strong>of</strong>enac at 2.1 Å Resolution: Evidence for an Induced Fit Model<br />

<strong>of</strong> Substrate Binding, Biochemistry 42, 9335-45.<br />

Whitacre, D. M., Monteiro, S. C. <strong>and</strong> Boxall, A. B. A. (2010), Occurrence <strong>and</strong> fate <strong>of</strong><br />

human <strong>pharmaceuticals</strong> in the environment<br />

reviews <strong>of</strong> environmental contamination <strong>and</strong> toxicology, 202, pp. 53-154 (Springer New<br />

York,<br />

WHO Collaborating Centre for Drug Statistics Methodology (2012), The Anatomical<br />

Therapeutic Chemical (ATC),<br />

Wickramasinghe, R. H. <strong>and</strong> Villee, C. A. (1975), Early role during chemical evolution for<br />

cytochrome P450 in oxygen detoxification, Nature 256, 509-11.<br />

Wiegel, S.et al. (2004), Pharmaceuticals in the river Elbe <strong>and</strong> its tributaries, Chemosphere<br />

57, 107-26.<br />

Williams, D. G., Patel, A. <strong>and</strong> Howard, R. F. (2002), Pharmacogenetics <strong>of</strong> codeine<br />

metabolism in an urban population <strong>of</strong> children <strong>and</strong> its implications for analgesic<br />

reliability, Br J Anaesth 89, 839-45.<br />

Williams, P. A., Cosme, J., Sridhar, V., Johnson, E. F. <strong>and</strong> McRee, D. E. (2000),<br />

Mammalian Microsomal Cytochrome P450 Monooxygenase: Structural<br />

Adaptations for Membrane Binding <strong>and</strong> Functional Diversity, Mol Cell 5, 121-31.<br />

Winter, H. R. <strong>and</strong> Unadkat, J. D. (2005), IDENTIFICATION OF CYTOCHROME P450<br />

AND ARYLAMINE N-ACETYLTRANSFERASE ISOFORMS INVOLVED IN<br />

SULFADIAZINE METABOLISM, Drug Metab Dispos 33, 969-76.<br />

Wolfenden, B. S. <strong>and</strong> Willson, R. L. (1982), Radical-cations as reference chromogens in<br />

kinetic studies <strong>of</strong> ono-electron transfer reactions: pulse radiolysis studies <strong>of</strong> 2,2'-<br />

azinobis-(3-ethylbenzthiazoline-6-sulphonate), J Chem Soc, Perkin Transactions 2<br />

805-12.<br />

World Health Organization (2011), WHO Model List <strong>of</strong> Essential Medicines, 17th list,<br />

World Health Organization (2012), World Health Organization. WHO Model Lists <strong>of</strong><br />

Essential Medicines,<br />

Wrighton, S. A. <strong>and</strong> Stevens, J. C. (1992), The Human Hepatic Cytochromes P450<br />

Involved in Drug Metabolism, CRC Cr Rev Toxicol 22, 1-21.<br />

Yarman, A.et al. (2012), The aromatic peroxygenase from Marasmius rotula—a new<br />

enzyme for biosensor applications, Anal Bioanal Chem 402, 405-12.<br />

Yu, J., Chang, P.-K., Ehrlich, K. C., Cary, J. W., Montalbano, B., Dyer, J. M., Bhatnagar,<br />

D. <strong>and</strong> Clevel<strong>and</strong>, T. E. (1998), Characterization <strong>of</strong> the Critical Amino Acids <strong>of</strong> an<br />

Aspergillus parasiticus Cytochrome P-450 Monooxygenase Encoded <strong>by</strong> ordA That<br />

141


REFERENCES<br />

Is Involved in the Biosynthesis <strong>of</strong> Aflatoxins B1, G1, B2, <strong>and</strong> G2, Appl Environ<br />

Microbiol 64, 4834-41.<br />

Yun, C. H., Miller, G. P. <strong>and</strong> Guengerich, F. P. (2000), Rate-determining steps in<br />

phenacetin oxidations <strong>by</strong> human cytochrome P450 1A2 <strong>and</strong> selected mutants,<br />

Biochemistry 39, 11319-29.<br />

Zimatkin, S. M. <strong>and</strong> Deitrich, R. A. (1997), Ethanol metabolism in the brain, Addict Bio 2,<br />

387-400.<br />

142


APPENDIX<br />

6 Appendix<br />

6.1 Aromatic ring hydroxylation<br />

Appendix Figure 1 HPLC-elution pr<strong>of</strong>iles <strong>of</strong> reaction mixtures showing the formation <strong>of</strong> products<br />

after conversion <strong>of</strong> propranolol (0.5 mM) <strong>by</strong> the aromatic peroxygenases (2 Units mL -1 ) <strong>of</strong> A.<br />

aegerita (AaeAPO) <strong>and</strong> C.radians (CraAPO).<br />

Appendix Figure 2 HPLC-elution pr<strong>of</strong>iles <strong>of</strong> reaction mixtures showing the formation <strong>of</strong> products<br />

after conversion <strong>of</strong> dicl<strong>of</strong>enac (0.5 mM) <strong>by</strong> the aromatic peroxygenase (2 Units mL -1 ) <strong>of</strong> A.<br />

aegerita. The insets show the UV-Vis-spectra <strong>of</strong> products <strong>and</strong> the known structures: (I) -<br />

dicl<strong>of</strong>enac; (II) – 4'-hydroxydicl<strong>of</strong>enac; (III) – dicl<strong>of</strong>enac-1',4'-quinone imine.<br />

143


APPENDIX<br />

6.2 Oxidation <strong>of</strong> alkyl side chains<br />

Appendix Figure 3 HPLC-elution pr<strong>of</strong>iles <strong>of</strong> reaction mixtures showing the formation <strong>of</strong> products<br />

after the conversion <strong>of</strong> tolbutamide (0.5 mM) <strong>by</strong> the aromatic peroxygenase (2 U mL -1 ) <strong>of</strong> A.<br />

aegerita (AaeAPO) with <strong>and</strong> without ascorbic acid. The insets show mass- <strong>and</strong> UV-Vis-spectra <strong>of</strong><br />

tolbutamide <strong>and</strong> its products with structures. I – tolbutamide; II – 4-ketotolbutamide; III – 4-<br />

hydroxytolbutamide.<br />

6.3 O-Dealkylation<br />

Appendix Figure 4 HPLC-elution pr<strong>of</strong>iles <strong>of</strong> reaction mixtures showing the formation <strong>of</strong> products<br />

after conversion <strong>of</strong> naproxen (0.5 mM) <strong>by</strong> the aromatic peroxygenase (2 U mL -1 ) <strong>of</strong> A. aegerita<br />

(AaeAPO) with ascorbic acid. The insets show the mass spectra <strong>of</strong> naproxen (II) <strong>and</strong> O-<br />

desmethylnaproxen (I).<br />

144


APPENDIX<br />

6.4 N-dealkylation<br />

Appendix Figure 5 HPLC-elution pr<strong>of</strong>iles <strong>of</strong> reaction mixtures showing the formation <strong>of</strong> products<br />

after conversion <strong>of</strong> sildenafil (0.5 mM) <strong>by</strong> AaeAPO, MroAPO <strong>and</strong> a recombinant peroxygenase<br />

from Novozymes (2 U mL -1 ). The insets show the mass spectra <strong>of</strong> sildenafil (I), N-<br />

desmethylsildenafil (II) <strong>and</strong> proposed <strong>other</strong> products (III) <strong>and</strong> (IV).<br />

6.5 Oxidation <strong>of</strong> steroids<br />

Appendix Figure 6 HPLC-elution pr<strong>of</strong>iles <strong>of</strong> reaction mixtures showing the formation <strong>of</strong> products<br />

after conversion <strong>of</strong> pregnenolone (0.5 mM) <strong>by</strong> the aromatic peroxygenase (2 U mL -1 ) <strong>of</strong> M. rotula.<br />

The insets show the mass spectra <strong>of</strong> two products <strong>and</strong> known structures: (I)–pregnenolone; (II)–<br />

progesterone; (III), (IV) & (V)–not identified products.<br />

145


APPENDIX<br />

6.6 O-Demethylenation<br />

Appendix Figure 7 HPLC-elution<br />

pr<strong>of</strong>iles <strong>of</strong> reaction mixtures showing<br />

the formation <strong>of</strong> 4-nitrocatechol after<br />

the oxidation <strong>of</strong> NBD (0.5 mM) <strong>by</strong><br />

the aromatic peroxygenases (2 Units<br />

mL -1 ) <strong>of</strong> C. verticillata (I), C. radians<br />

(II) <strong>and</strong> A. aegerita (III). The<br />

oxidation <strong>of</strong> NBD <strong>by</strong> M. rotula<br />

peroxygenase <strong>and</strong> recombinant C.<br />

cinerea peroxygenase was also tested<br />

<strong>and</strong> gave 4-nitrocatechol as major<br />

metabolite as well (data not shown).<br />

Appendix Figure 8 HPLC-elution pr<strong>of</strong>iles <strong>of</strong> reaction mixtures showing the formation <strong>of</strong> products<br />

after conversion <strong>of</strong> bromobenzodioxole (0.5 mM) <strong>by</strong> the aromatic peroxygenase (2 U mL -1 ) <strong>of</strong> A.<br />

aegerita. The oxidation <strong>of</strong> chlorobenzodioxole gave a similar elution pr<strong>of</strong>ile <strong>and</strong> corresponding<br />

products (data not shown). The insets show mass spectra <strong>of</strong> products <strong>and</strong> the known structures: (I)<br />

– bromobenzodioxole; (II) – bromocatechol; (III) – not identified product; (IV) –<br />

trihydroxybromobenzene (bromopyrrogallol).<br />

146


APPENDIX<br />

6.7 Spectrophotometric nitrobenzodioxole assay<br />

Appendix Figure 9 Determination <strong>of</strong> the<br />

extinction coefficient <strong>of</strong> 4-nitrocatechol at<br />

pH >12. UV-Vis spectra <strong>of</strong> different<br />

concentrations <strong>of</strong> 4-nitocatechol (12.5, 25,<br />

50, 62.5, 100, 125 µM) dissolved in<br />

potassium phosphate/NaOH (50 mM/100<br />

mM) were recorded <strong>and</strong> the extinction<br />

coefficient calculated for the maximum at<br />

514 nm. The inset shows the linear<br />

relationship between the concentration <strong>of</strong> 4-<br />

nitrocatechol <strong>and</strong> the absorbance at 514 nm.<br />

6.8 Stoichiometrical analyses<br />

Appendix Table 1 Stoichiometry <strong>of</strong> nitrobenzodioxole oxidation <strong>by</strong> A. aegerita<br />

peroxygenase. 1<br />

H 2 O 2 added catechol produced catechol/H 2 O 2<br />

23.44<br />

46.88<br />

70.3<br />

93.75<br />

187.5<br />

µM<br />

15.1<br />

29.21<br />

39.45<br />

49.87<br />

80.50<br />

0.64 ± 0.01<br />

0.62 ± 0.03<br />

0.56 ± 0.04<br />

0.53 ± 0.01<br />

0.43 ± 0.03<br />

1 The initial nitrobenzodioxole concentration was 0.25 mM.<br />

147


APPENDIX<br />

6.9 Experiment with propranolol at an environmentally relevant<br />

concentrations<br />

Ausgangslösung<br />

1800 ml bi-destiliertes H 2 O<br />

inclusive 2 ml 5 µM Propranolol<br />

+ 900 ml Ausgangslösung<br />

+ 100 ml bi-destiliertes H 2 O<br />

= 1000 ml mit 5 nM Propranolol<br />

1. Ansatz<br />

+ 900 ml Ausgangslösung<br />

+ 50 ml (inclusive 100 Units AaeApo)<br />

+ 50 ml (inclusive 200 µl 30% H 2 O 2<br />

= 1000 ml mit 5nM Propranolol<br />

100 U/l AaeApo und 2 mM H 2 O 2<br />

2. Umsatz<br />

18 h<br />

Inkubationszeit<br />

zur Vergleichbarkeit mit der<br />

Probelösung<br />

3. Ultrafiltration<br />

zur Enzymabtrennung<br />

mit 10 molarer Natronlauge<br />

auf pH 12<br />

4. pH-Wert Einstellung<br />

mit 10 molarer Natronlauge<br />

auf pH 12<br />

mit 25 ml Dichlormethan im<br />

Scheidetrichter<br />

5. Extraktion<br />

mit 25 ml Dichlormethan<br />

im Scheidetrichter<br />

Dichlormethan mittels Turbo-Vap auf<br />

1 ml aufkonzentrieren<br />

6. Aufkonzentration<br />

Dichlormethan mittels Turbo-Vap<br />

auf 1 ml aufkonzentrieren<br />

Dichlormethan mit 2 ml Acetonitril<br />

mischen und auf 1 ml<br />

aufkonzentrieren<br />

7. Extraktion durch<br />

Konzentration<br />

Dichlormethan mit 2 ml Acetonitril<br />

mischen und auf 1 ml<br />

aufkonzentrieren<br />

Acetonitril mit<br />

2 ml bi-destiliertem H 2 O mischen und<br />

auf 1 ml aufkonzentrieren<br />

8. Extraktion durch<br />

Konzentration<br />

Acetonitril mit 2 ml bi-destiliertem<br />

H 2 O mischen und auf 1 ml<br />

aufkonzentrieren<br />

HPLC<br />

„Luna C18“-Säule<br />

9. Analyse<br />

HPLC<br />

„Luna C18“-Säule<br />

9. Vergleich<br />

Appendix Figure 10 Process flow diagram <strong>of</strong> the experiment with AaeAPO <strong>and</strong> an<br />

environmentally relevant concentration (5 nM) <strong>of</strong> propranolol (in German). Figure adapted<br />

from Sylvia Gleissner (Master thesis supervised at the <strong>IHI</strong> <strong>Zittau</strong>)<br />

148


APPENDIX<br />

Appendix Figure 11 Examplary 3D-structures (stick models) <strong>of</strong> <strong>pharmaceuticals</strong>: sildenafil<br />

(A); dicl<strong>of</strong>enac (B) <strong>and</strong> dextromethorphan (C).<br />

149


APPENDIX<br />

6.10 List <strong>of</strong> SCI publications<br />

1. Poraj-Kobielska, M., Kinne, M., Ullrich, R., Scheibner, K. & H<strong>of</strong>richter M. (2012).<br />

A spectrophotometric assay for the detection <strong>of</strong> <strong>fungal</strong> peroxygenases. Analytical<br />

Biochemistry, Vol. 421: 327-329. (doi: 10.1016/j.ab.2011.10.009) (impact factor: 3.0)<br />

2. Poraj-Kobielska, M., Kinne, M., Ullrich, R., Scheibner, K., Kayser, G., Hammel, K.<br />

E. <strong>and</strong> H<strong>of</strong>richter M. (2011). Preparation <strong>of</strong> human drug metabolites using <strong>fungal</strong><br />

peroxygenases. Biochemical Pharmacology, Vol. 82: 789-796 (doi:<br />

10.1016/j.bcp.2011.06.020) (impact factor: 4.7)<br />

3. Kinne, M., Poraj-Kobielska, M., Ullrich, R., Nousiainen, P.,Sipilä, J.,Scheibner, K.,<br />

Hammel, K. E., H<strong>of</strong>richter, M. (2011). Oxidative cleavage <strong>of</strong> non-phenolic β -O-4<br />

lignin model dimers <strong>by</strong> an extracellular aromatic peroxygenase. Holzforschung, Vol.<br />

65: 673–679. (impact factor: 1.8)<br />

4. Kinne, M., Poraj-Kobielska, M., Ralph, S. A., Ullrich, R., H<strong>of</strong>richter, M. <strong>and</strong><br />

Hammel, K. E. (2009). Oxidative cleavage <strong>of</strong> diverse ethers <strong>by</strong> an extracellular <strong>fungal</strong><br />

peroxygenase. The Journal <strong>of</strong> Biological Chemistry, Vol. 284: 29343-29349. (impact<br />

factor: 5.3)<br />

5. Kinne, M. Poraj-Kobielska, M., Ar<strong>and</strong>a, E., Ullrich, R., Hammel, K. E., Scheibner,<br />

K., & H<strong>of</strong>richter, M. (2009). Regioselective preparation <strong>of</strong> 5-Hydroxypropranolol <strong>and</strong><br />

4’-Hydroxydicl<strong>of</strong>enac with a <strong>fungal</strong> peroxygenase. Bioorganic & Medicinal<br />

Chemistry Letters, Vol. 19: 3085–3087. (impact factor: 2.5)<br />

150


ACKNOWLEDGEMENT<br />

Acknowlegement<br />

This dissertation could not have been written without the support <strong>of</strong> many people.<br />

Foremost, I want to express my special gratitude to pr<strong>of</strong>essor Martin H<strong>of</strong>richter who not<br />

only served as my supervisor but also encouraged <strong>and</strong> challenged me throughout<br />

this project, also for his persistence, <strong>and</strong> patience. He <strong>and</strong> my laboratory advisor<br />

Dr. Matthias Kinne guided me through the dissertation process, never accepting<br />

less than my best efforts. I thank them both.<br />

I would like to thank pr<strong>of</strong>essor Kenneth Hammel <strong>and</strong> pr<strong>of</strong>essor Korneliusz Miksch for<br />

willingness to be reviewer <strong>of</strong> my work.<br />

Special thanks again to Dr. Matthias Kinne for his guidance, assistance, patience <strong>and</strong><br />

advice regarding the selection <strong>and</strong> operation <strong>of</strong> research methods <strong>and</strong> ideas.<br />

I want to express my gratitude to Dr. René Ullrich, Martin Kluge <strong>and</strong> Marek Pecyna for<br />

theirs assistance with the analytical instrumentation <strong>and</strong> fruitful discussions.<br />

I would like to thank my laboratory colleagues Anna Karutz, Sylvia Gleißner, Stefanie<br />

Hintersatz, Małgorzata Poraj-Kobielska, Natalia Lemańska-Malinowska <strong>and</strong> the<br />

laboratory technicians Ulrike Schneider <strong>and</strong> Monika Br<strong>and</strong>t for their technical <strong>and</strong><br />

scientific assistance.<br />

I want also express my gratitude to Dr. Gernot Kayser for fruitful discussions <strong>and</strong> help <strong>by</strong><br />

organize <strong>of</strong> financial support for this project.<br />

Many thanks to whole Laboratory team for pleasant working environment <strong>and</strong> the nice<br />

time together.<br />

Lastly, I <strong>of</strong>fer my regards <strong>and</strong> thanks to all <strong>of</strong> those who supported me in any respect<br />

during the completion <strong>of</strong> the project, my family <strong>and</strong> friends for their<br />

encouragement <strong>and</strong> underst<strong>and</strong>ing.


Versicherung an Eides statt<br />

Hiermit versichere ich, dass ich die vorliegende Arbeit ohne unzulässige Hilfe Dritter und<br />

ohne Benutzung <strong>and</strong>erer als der angegebenen Hilfsmittel angefertigt habe; die aus fremden<br />

Quellen direkt oder indirekt übernommenen Gedanken sind als solche kenntlich gemacht.<br />

Weitere Personen waren an der Abfassung der vorliegenden Arbeit nicht beteiligt. Die Hilfe<br />

eines Promotionsberaters habe ich nicht in Anspruch genommen. Weitere Personen haben von<br />

mir keine geldwerten Leistungen für Arbeit erhalten, die nicht als solche kenntlich gemacht<br />

worden sind.<br />

Die Arbeit wurde bisher weder im Inl<strong>and</strong> noch im Ausl<strong>and</strong> in gleicher oder ähnlicher Form<br />

einer <strong>and</strong>eren Prüfungsbehörde vorgelegt.<br />

<strong>Zittau</strong>, den 5. Oktober 2012<br />

Marzena Poraj-Kobielska

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!