Download C 1 G

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Discovery and development of proton pump inhibitors wikipedia , lookup

Psychopharmacology wikipedia , lookup

Compounding wikipedia , lookup

Discovery and development of ACE inhibitors wikipedia , lookup

Discovery and development of neuraminidase inhibitors wikipedia , lookup

Neuropsychopharmacology wikipedia , lookup

Discovery and development of non-nucleoside reverse-transcriptase inhibitors wikipedia , lookup

Discovery and development of cephalosporins wikipedia , lookup

Neuropharmacology wikipedia , lookup

Theralizumab wikipedia , lookup

Bad Pharma wikipedia , lookup

Medication wikipedia , lookup

Pharmaceutical industry wikipedia , lookup

Prescription costs wikipedia , lookup

Prescription drug prices in the United States wikipedia , lookup

Drug design wikipedia , lookup

Pharmacokinetics wikipedia , lookup

Pharmacogenomics wikipedia , lookup

Drug interaction wikipedia , lookup

Drug discovery wikipedia , lookup

Pharmacognosy wikipedia , lookup

Transcript
CHAPTER 1
GENERAL INTRODUCTION
Chapter 1
General Introduction
Introduction
1.
Role of drug metabolism in drug development
4
1.1. Metabolites in safety testing (MIST)
1.2. Enzyme involved in drug metabolism
1.3. Biosynthesis of reactive metabolites
1.4. Biological screening for active metabolites
2.
Microbial conversion of steroid compounds
12
3.
Cytochrome P450s
15
3.1. Human P450s
3.2. Structural organization of P450s
3.3. The catalytic cycle
4.
Cytochrome P450 BM3
20
4.1. Engineering Cytochrome P450 BM3
4.2. P450 BM3 in biocatalysis
4.3. Key residues in the active site of P450 BM3
5.
Spin lattice relaxation NMR
27
6.
Scope and objective of this thesis
29
7.
Outline of this thesis
30
Chapter 1
General Introduction
1.
Role of drug metabolism in drug discovery and development
Most xenobiotic substances such as drugs undergo an array of biotransformation
reactions after internal exposure to humans. Drug metabolism by the host system is one
of the most important determinants of the pharmacokinetic profile of a drug.
Biotransformation of drugs can have different effects, such as the formation of chemically
stable metabolites, which are devoid of pharmacological or toxicological activities, or the
generation of short lived chemically reactive metabolites, which can lead to toxicological
side effects (1). Furthermore, drug metabolism can also lead to formation of
pharmacologically active metabolites which might contribute significantly to drug action in
vivo, or may be responsible for pharmacologically-based adverse drug reactions.
A relatively novel approach in modern drug development is the use of pharmacologically
active metabolites as potential resources for drug discovery and development. There are
several advantages for screening drug candidates for active metabolites during drug
discovery. For example drug metabolites can show superior properties compared to the
lead itself, such as improved pharmacodynamics, improved pharmacokinetics, lower
probability of drug-drug interactions, less variable pharmacokinetics and/or
pharmacodynamics, improved overall safety profile and improved physicochemical
properties (e.g. solubility) (1).
Examples of active metabolites of previously marketed drugs that have been developed
into drugs are listed in Table 1.
Table 1. Pharmacologically active metabolites developed as new drugs (2)
Parent drugs
Metabolite drugs
Biotransformation
Brand name
Allopurinol
Oxypurinol
Oxidation of xanthine
Oxyprim
Amitriptyline
Nortryptyline
N-Demethylation
Aventyl
Bromhexine
Ambroxol
N-Demethylation
& Mucosovan
Hydroxylation
Diazepam
Oxazepam
N-demethylation
& Serax
Hydroxylation
Etretinate
Acitretin
Deesterification
Soriatane
Hydroxyzine
Cetirizine
Carboxylation
Zyrtec
Imipramine
Desimipramine
N-Demethylation
Norpramin
Loratadine
Desloratadine
Descarboethoxylation
Clarinex
Loxapine
Amoxapine
N-Demethylation
Asendin
Phenacetin
Acetaminophen
O-Deethylation
Tylenol
Terfenadine
Fexofenadine
Carboxylation
Allegra
Thioridazine
Mesoridazine
S-Oxidation
Serentil
Chapter 1
General Introduction
For example, fexofenadine, the primary metabolite of terfenadine was developed as antihistamine agent with selective peripheral histamine H1-receptor antagonist activity.
Fexofenadine is a safer drug than terfenadine because it does not has the ability to cause
cardiotoxicity by hERG inhibition.
In some instances, metabolic transformation can also produce reactive or toxic
intermediates or metabolites, with potential toxicological implications (3). Hence, a good
understanding of the metabolism of a new chemical entity is needed early in the drug
discovery process.
1.1. Metabolites in safety testing (MIST)
Drug metabolism and pharmacokinetics are crucial factors for successful drug
development. Inappropriate pharmacokinetics previously was one of the main reasons of
attrition in drug development. Although inappropriate pharmacokinetics has become a
less important reason for attrition, less progress has been made in decreasing the attrition
due to drug toxicity. The preparative synthesis of drug metabolites is currently of primary
importance in industry in order to assess potential toxicity, drug-drug interactions and to
examine metabolic pathways (4). In fact, testing the toxicities and biological activities of
human metabolites of drugs is important in drug development to assure effectiveness and
safety (5).
The importance of metabolite identification and quantification is stressed by the
guidelines published in 2008 by the US food and drug administration (FDA). The so-called
Metabolites in Safety Testing (MIST) guidelines describe the studies recommended to
perform on metabolites having more than 10% systemic exposure of the parent drug to
support human safety. A decision diagram of this is depicted in Figure 1.
These guidelines challenged the development of biocatalytic systems for the generation of
large amounts of drug metabolites as well as analytical technologies dealing with their
quantification and identification of drug metabolites in an efficient and intelligent way (7).
Chapter 1
General Introduction
Figure 1. Decision tree flow diagram of the MIST guidelines as determined by the FDA,
based on (6).
1.2. Enzymes involved in drug metabolism
Drug metabolism is typically divided in two phases. Phase I metabolism includes many
types of functionalization reactions, e.g. oxidation, reduction, hydrolysis, hydration and
dehalogenation reactions (9). Cytochrome P450s represent the most important class of
enzymes involved in phase I metabolism, being involved in 75-80% of metabolism of
marketed drugs, Figure 2. Other phase I enzymes include monoamine oxidases, flavincontaining oxygenases, amidases and esterases (9). Phase II metabolism involves
conjugation of polar groups (e.g. glucuronic acid, sulfate, and amino acids) to drugs and/or
phase I drug metabolites to further increase their hydrophilicity and facilitate excretion.
These biotransformation reactions involve glucuronidation, sulfation, GSH conjugation,
acetylation, amino acid conjugation and methylation reaction (10). Phase II enzymes
include UDP-glucuronosyltransferase (UGTs), sulfotransferases (STs), N-acetyl transferases
(NATs), methyl transferases and glutathione S-transferases (GSTs) (10), Figure 2.
Almost all phase I and phase II enzymes can catalyze the formation of stable metabolites,
as well as chemically reactive intermediates that may lead to the generation of adverse
drug reactions (11).
Chapter 1
General Introduction
Figure 2. Phase I and phase II metabolism of currently marketed drugs (adapted from
(8)). The percentage of phase I and phase II metabolism of drugs that each enzyme
contributes is estimated by the relative size of each section of the corresponding chart.
ADH, alcohol dehydrogenase; ALDH, aldehyde dehydrogenase; DPD, dihydropyrimidine
dehydrogenase; NQO, NADPH quinine oxidoreductase.
COMT, catechol Omethyltransferase; GST, glutathione S-transferase; HMT, histamine methyltransferase;
NAT, N-acetyltransferase; TPMT, thiopurine methyltransferase; UGTs, uridine 5’triphosphate glucuronosyltransferases.
1.3. Role of reactive metabolites in adverse drug reactions
Adverse drug reactions (ADRs), as defined by the World Health Organization, are noxious,
unintended and undesirable effects of a drug, which occurs at therapeutical doses used in
humans and which require hospitalization of patients to recover (11). Although much
effort has been spent to the development of safer drugs, ADRs still constitute a major
reason for attrition in drug development and withdrawal or restriction of marketed
products. Therefore, prevention of ADRs constitutes a major challenge for the
pharmaceutical industry (3).
Approximate 75% of ADR in patients are classified as Type A ADR which involve lifethreatening exaggerated pharmacological activity (on-target and/or off-target) due to
unanticipated increased plasma concentrations of the parent drug. This class of ADR often
results from drug-drug interactions or genetic deficiency of an enzyme which is involved in
the major pathway of metabolism. For the other classes of ADR formation of reactive drug
metabolites are considered to play a crucial role (12).
The main difficulty for pharmaceutical industry is the fact that ADRs cannot always be
predicted in preclinical animal studies. In particular idiosyncratic drug reactions (IDRs;
Type B ADR) are difficult to predict as they normally have a very low incidence, have a
Chapter 1
General Introduction
delayed time of onset, do not necessarily show classic dose-response relationships and are
not predictable from the known pharmacology of the drug. No predictive animal model is
currently available and IDRs therefore remain poorly understood.
Figure 3 shows a general scheme depicting the role of reactive metabolites in the
development of toxicity. Drugs bioactivation can be catalyzed by both phase I and phase II
enzymes.
Different types of reactive metabolites can be distinguished such as electrophiles, radicals
and reactive oxygen species. Electrophilic metabolites can react with nucleophilic sites in
proteins and/or DNA. Free radicals possess unpaired electrons and can abstract hydrogen
atoms from polyunsaturated fatty acids in membranes which will initiate the destructive
process of lipid peroxidation, and can disrupt the cellular redox-state leading to oxidative
stress and subsequent toxicity (12).
Figure 3. Scheme depicting the role of drug metabolism in determining the toxic
outcome in ADRs.
Table 2 gives examples of drugs whose toxicities are attributed to covalent binding of
electrophilic metabolites to proteins. Bioactivation of xenobiotics which contain certain
functional groups such as tertiary amine, furan ring or acetylene group, can also cause
mechanism-based inactivation of P450, leading to adverse clinical drug-drug interactions
(13-14).
Chapter 1
General Introduction
Table 2. Examples of drugs and reactive metabolites possibly involved in ADRs (12-15).
Drugs
Reactive intermediate
Toxicity
Acetaminophen
Quinone-imine
Hepatotoxicity
Carbamazepine
2-Hydroxy, Quinone-imine
Agranulocytosis, Aplastic
anemia
Clarithromycin
Nitroalkane
Hypersensitivity
Clozapine
Nitrenium ion
Agranulocytosis
Dapsone
Hydroxulamine, Nitroso
Hemolysis, hypersensitivity
Diclofenac
Acylglucuronide, Benzoquinone imine
Hepatotoxicity
Halothane
Trifluoroacetyl
Hepatitis
Indomethacin
Iminoquinone
Hypersensitivity,
Agranulocytosis,
Hepatotoxicity
Isonizaid
Isonicotinic acid, acetylating species
Hypersensitivity
Phenacetin
p-Nitrosophenetole
Hepatotoxicity
Procainamide
Hydroxylamine, Nitroso
Agranulocytosis
Tacrine
7-OH-tacrine
Hepatotoxicity
Tamoxifen
N-oxide, N-oxide-epoxide
Carcinogenesis
Ticlodipine
Keto, S-oxide
Agranulocytosis, aplastic
anemia
Tielinic acid
Thiopene S-oxide
Hypersensitivity, hepatitis
Troglitazone
Conjugates, Benzoquinone, Quinone
Hepatotoxicity
epoxide
Valproic acid
Acyl glucuronides
Hypersensitivity,
Hepatotoxicity
Detection of reactive metabolites is difficult due to the chemical reactivity of the shortlived intermediates and to the usually low amounts present in incubations. The most
common way to screen for formation of reactive metabolites is to trap them with model
nucleophiles such as GSH and to analyze the formed adducts with spectroscopic
techniques. The endogenous tripeptide GSH can trap different types of RIs including
quinones, quinoneimines, iminoquinone, methides, epoxides, areneoxides and nitrenium
ions (15). The corresponding GSH adducts are typically analyzed by liquid-chromatography
mass spectrometry (LC-MS). However, a major disadvantage is that structural information
of the adducts obtained by LC-MS/MS experiments is limited, as it may be insufficient in
identifying the exact position of oxidation, to differentiate stereoisomers, or to provide
the exact structure of the metabolite (16). Complementing MS data with Nuclear
Magnetic Resonance (NMR) experiments is required for complete structural elucidation of
drug metabolites and/or resulting adducts. A limitation of NMR, however, is its relative
insensitivity, as it requires several milligrams of pure metabolite.
Chapter 1
General Introduction
The generation of pure metabolites can be achieved by organic synthesis,
electrochemical oxidation of the parent drug or by biocatalysis using appropriate enzymes.
The synthesis by organic chemistry is often complicated by the lack of appropriate
synthetic routes for specific metabolites and may yield only low amounts of the desired
product that has to be purified subsequently.
Electrochemical oxidation has been used to generate drug metabolites and has the
advantage that it can be easily scaled up. However, with this technique not all the relevant
enzymatic metabolites can be formed. Electrochemistry has been previously used to
generate GSH adducts of clozapine (17) and troglitazone (18).
Biocatalysis by human or bacterial P450s can also be used for the scaling up of metabolite
production (19). For example, human relevant diclofenac metabolites were synthesized by
microbial fermentation in large amount (up to 170 mg) and their structures were
elucidated by MS and NMR (20).
More recently the use of highly active bacterial P450 BM3 mutants for the generation of
reactive metabolites of drugs has been explored. Damsten et al. showed that BM3
mutants could be used for the generation of human relevant reactive metabolites of
clozapine, acetominophen, diclofenac (21), and trimethoprim (22). Boerma et al. showed
that P450 BM3 mutants with high capacity to activate drugs (clozapine, acetaminophen
and troglitazone) into relevant reactive metabolites can be employed to produce protein
adducts to study the nucleophilic selectivity of highly reactive electrophiles (23). Dragovic
et al. applied highly active BM3 mutant as bioactivation system to study the role of human
GSTs in the inactivation of reactive intermediates of clozapine (24).
1.4. Biological screening for active metabolites
To assess the biological activity of drug metabolites there are several approaches
available. The traditional approach is to isolate and purify all the metabolites from in vitro
incubations with subcellular fractions such as liver microsomes or intact cellular systems
(e.g. epatocytes) containing a full complement of drug metabolizing enzymes.
An alternative approach is to use bioassay-guided methods where biological samples
containing biotransformation products are first evaluated for their pharmacological
activity prior to isolation of the metabolites. The bioassay methods may be based on the
assessment of the pharmacological activity using in vitro ligand binding (25, 26), cell based
assays (1), or in vivo pharmacological assays (1).
One of the major bottlenecks in in vitro assays is the lack of possibility to screen affinity or
activity of individual compounds in complex mixtures.
In general, two approaches can be used to measure the effect of individual
compounds in a mixture. One strategy is to combine HPLC with fractionation techniques
to enable the compounds in the mixture to be separated and allow screening of the
individual components using plate reader assays.
Chapter 1
General Introduction
The other available strategy is the on-line high resolution screening technique (HRS)
depicted in Figure 4 (27-29).
HRS enables the screening of individual compounds in complex mixtures by coupling a
separation technology, usually gradient HPLC, to post-column biochemical detection
assays on-line. This technique has already been successfully used to develop a number of
antibody- (30) and biotin- (31), receptor- (32), enzyme- (33) (34) (35), and MS-based (36)
screening assays. The HRS platforms are especially attractive when complex mixtures such
as metabolic incubations (37), combinatorial mixtures (38), or natural extracts (39) are
screened. In contrast to conventional high-throughput screening (HTS), isolation of
individual metabolites not showing affinity can be avoided. Thus, the laborious
development of high-yield synthesis and isolation methods can be limited to compounds
with the desired affinity.
Figure 4. Scheme of the online setup. The system enables the separation of mixtures and
subsequent parallel detection of enzyme binding and accurate mass. The samples are
injected into and separated by an HPLC system. The eluent is split between HR-MS and
enzyme binding detection. The fraction which enters the enzyme binding detection is
mixed with the target (e.g. p38) and the tracer and incubated after each mixing step,
24 s for the target–ligand interaction and 11 s for the target–tracer interaction. Finally,
the fluorescence is measured as readout of affinity towards the target. In parallel, the
second part of the eluent is analyzed by HR-MS delivering structural information to
identify the small molecules tested. Adapted from (28).
2.
Microbial conversion of steroid compounds
Chapter 1
General Introduction
Totally, about 300 approved steroid drugs are known to date, and this number tends to
grow (40). Steroid pharmaceuticals are ranked among the most marketed pharmaceuticals
and represent the second largest category next to antibiotics.
The research on steroid drugs started in 1950, with the discovery of the pharmacological
effects of cortisol and progesterone, two endogenous steroids, and with the identification
of the 11-hydroxylation activity of a Rhizopus species (41-42). Microbial steroid
transformation is a powerful tool for generation of novel steroidal drugs, as well as for
efficient production of steroid-containing active pharmaceutical ingredients (APIs) and key
intermediates.
Steroid hydroxylation is a key tool to develop steroid drugs with improved potency, longer
half-lives in the blood stream, simpler delivery methods and reduced side effects (43). The
therapeutic area of hydroxy-steroids is very wide: they are used as anti inflammatory,
immunosuppressive, progestational, diuretic, anabolic and contraceptive agents (41, 4447). Applications are found in the treatment of adrenal insufficiencies, in the prevention of
coronary heart disease, as anti-fungal agents, as anti-obesity agents, and in the inhibition
of HIV integrase, the prevention and treatment of infection by HIV and in the treatment of
declared AIDS (43).
The complex structure of the steroid molecule requires complicated, multi-step schemes
for the chemical synthesis of steroid compounds. The selective oxidation of an
unactivated, aliphatic C-H bond to an alcohol is a challenging problem in synthesis (48).
Chemical synthesis requires the use of reagents as pyridine, sulfur trioxide or selenium dioxide, whose disposal constitutes an environment issue (43). Moreover, the chemical
catalysts used often show very low regioselectivity.
Steroid hydroxylation by microorganisms represents a powerful alternative to chemical
synthesis. Current trends in microbial hydroxylation studies cover the search for novel
biocatalysts capable of performing reactions targeting specific positions in the steroid
molecule that are particularly important for the pharmaceutical industry (7α, 9α, 11α,
11β, 16α, 17α), production of novel hydroxysteroids with potent therapeutic properties,
and the construction of novel biocatalysts by genetic engineering (40).
Examples are given in Table 3.
Cytochrome P450s play a very important role in endogenous steroid metabolism.
Testosterone hydroxylation in particular has been studied extensively with human liver
microsomal P450s. For example, CYP3A4, the most abundant P450 in human liver and
small intestine, catalyzes the monohydroxylation at 10 different positions of testosterone
(8). Hence testosterone hydroxylation patterns have been utilized as probes for the
presence and characterization of P450s (49) (50).
Hydroxylation of testosterone at 1-, 2-, 2-, 6-, 6-, 7-, 7-, 15-, 15-, 16-, 16and 17-positions has been reported in human and rat liver microsomes (Figure 5).
Table 3. Examples of steroid hydroxylation in microrganisms
Substrate
Product
Microrganism
Ref.
Chapter 1
7-Hydroxylation
Testosterone
Pregnenolone
Epiandrosterone
7-Hydroxulation
DHEA
7-15Dihydroxylation
DHEA
General Introduction
7-OH-testosterone
7-OH-pregnenolone
7-OH-epiandrostenone
Botrytis cinerea AM235
Fusarium oxysporum var.cubense
Mortierella isabellina AM212
(49)
7- and 7-derivatives
7-OH DHEA
M.racemosus
Botryodiplodia malorum CBS 134.50
(51)
(52)
7-15-diOH-DHEA (3,7,15- C.lini CBS 112.21
triOH-androst-5-ene-17-one)
7-15Dihydroxylation
Corynespora cassiicola CBS 161.60
17-OH-progesterone 8-OH-derivative, along with
15-hydroxy derivative
9-hydroxylation
Testosterone
9-OH-AD, 9-OH-testosterone R.equi ATCC 14887
Progesterone
9-OH-progesterone
Expression of 9-hydroxylase
from M.smegmatis in E.coli BL21
Androsterone
9-OH-adrenosterone, 9-OH- Cunninghamella elegans TSY-0865
11-keto-testosterone
11-Hydroxylation
Testosterone
Rhizopus stolonifer
11-OH-testosterone
ATCC 10404, Fusarium lini
NRRL 68751
11-hydroxylation
Progesterone
T.harzianum, T.hamatum
11-OH-progesterone
14-Hydroxylation
Testosterone
A.wentii MRC 200316
14-OH-testosterone
15-Hydroxylation
Testosterone
F.oxysporum var. cubense
15-OH testosterone
Progesterone
15-OH progesterone
15-Hydroxylation
Testosterone
Solvent tolerant
15-OH-testosterone
Pseudomonas putida S12
expressing B.megaterium
steroid hydroxyase CYP106A2
16-Hydroxylation
Testosterone
2,16,17-triOH-androst-4-en- Whetzelinia sclerotiorum
ATCC 18687
3-one
17-Hydroxylation
Progesterone
B.sphaericus ATCC 245,
17-OH-progesterone
B.sphaericus ATCC 7063,
B.sphaericus ATCC 13805
(50)
(52)
(53)
(54)
(55)
(56)
(57)
(58)
(59)
(60)
(61)
(62)
(63)
Chapter 1
General Introduction
Figure 5. Hydroxylation of Testosterone by human P450s (adapted from (51))
Figure 6. Testosterone hydroxylation by bacterial P450s (adapted from (40)).
Chapter 1
General Introduction
Hydroxylation of Testosterone by bacterial P450s has also been extensively studied (Figure
6) (51). Profiling this hydroxylation provides useful information for characterizing bacterial
P450s as monogenic traits and for comparing bacterial P450s with human P450s.
The features and functionality of steroid hydoxylases have been recently reviewed by, e.g.
Bernhardt (2006) (52), Hannemann et al (2007) (53), Novikova et al. (2009) (54) and
Kristan and Lanisnik-Rizner (2012) (55).
3.
Cytochrome P450s
Cytochrome P450s (P450s or CYPs) constitute a family of monoxygenases involved in the
biotransformation of drugs, xenobiotics, alkanes, terpenes, and aromatic compounds, in
the metabolism of chemical carcinogens, the biosynthesis of physiologically relevant
compounds as steroids, fatty acids, eicosanoids, fat-soluble vitamins, bile acids and in the
degradation of pesticides (52).
They are able to catalyze a number of difficult oxidative reactions, such as C-H bond
hydroxylation, N-dealkylation, N-hydroxylation, O-dealkylation, S-oxidation and
epoxidation of numerous endogenous and exogenous compounds (3). An overview of the
different reactions catalyzed by P450s is listed in Table 3.
The broad substrate acceptance and this huge diversity in catalyzed reactions make the
class of P450 very promising biocatalysts in industrial processes. In fact, in recent years
there has been an increasing interest in the application of CYP biocatalysts for the
industrial synthesis of bulk chemicals, pharmaceuticals, agrochemicals, and food
ingredients, especially when a high grade of stereo and regioselectivity is required (57).
Moreover P450s are of great interest in drug metabolism. They are ubiquitous enzymes
often involved in the oxidation of drugs resulting in the generation of metabolites that are
more easily excreted, thus modulating the toxicity of these compounds (57).
3.1. Human P450s
The human genome encodes for 57 functional P450s (58) of which approximately a dozen
are involved in drug metabolism. The P450s mostly involved in metabolism of drugs in
humans are listed in Table 5, where the main reaction and substrate specificities are
indicated. Genetic polymorphisms that may alter the enzyme activity have been reported
for a number of P450s (59). While for the P450s 1A2, 2A6, 2B6 and 2C8 the clinical effects
due to these polymorphisms are minor, major effects are observed for 2C9, 2C19, and 2D6
(60). Polymorphisms in these isoenzymes affect metabolism of 20% to 30% of the clinically
used drugs and this has become a major issue in the discovery and development of drugs
and other NCEs (61).
Table 4. List of different reactions catalyzed by P450s (56)
Reaction type
Example
Chapter 1
General Introduction
Aliphatic
hydroxylation
Aromatic
hydroxylation
Epoxidation
Heteroatom
dealkylation
Alkyne
oxygenation
Heteroatom
oxygenation
Aromatic
epoxidation and
NIH-shift
Dehalogenation
Dehydrogenation
Reduction
Cleavage of esters
Table 5.Substrate specificity of human CYPs
CYP
CYP1A2
Substrate properties
Drug/Substrates
Planar, polyaromatic
compounds
Clozapine
Phenacetine
Olanzepine
Imipramine Propranolol
Theophylline
Reaction
Oxidation
O-deethylation
Oxidation
Demethylation
4-hydroxylation
Chapter 1
3.2.
Structural organization of P450s
General Introduction
Chapter 1
General Introduction
All P450s are heme-containing proteins containing an iron atom at the center of the heme
which is also complexed by a cysteine residue as axial ligand. Based on their redox partner,
P450s can be classified in different classes (Figure 7):
Figure 7. Structural organization of P450s: (a) Class I (mostly bacterial P450s), (b) Class II
(mostly microsomal P450s) and (c) P450 BM3. (Adapted from (62))
Class I P450s (a) require two redox proteins: a small redox 2Fe-2S iron-sulfur ferredoxin
and a FAD or FMN containing reductase. Mitochondrial P450s and most soluble bacterial
P450s belong to this class (63).
In class II (b), electrons are transferred via only one reductase having a FAD and a FMN
domain. Microsomal P450s that are attached to the endoplasmatic reticulum belong to
this class (63).
Most bacterial P450s belong to class I. However, there are exceptions: in 1981
Cytochrome P450 BM3 (CYP102A1) from Bacillus megaterium was identified. P450 BM3
(c) is soluble and uses a Class II redox system. This soluble bacterial enzyme contains the
P450 monoxygenase domain fused to the electron transfer flavin mononucleotide
(FMN)/flavin adenine dinucleotide (FAD) reductase domain in a single polypeptide (63). In
contrast, mammalian CYPs require additional redox partners as cytochrome P450
reductase (CPR) and often cytocrome b5 and lipids (64, 65).
3.3. The catalytic cycle
The general catalytic cycle for substrate hydroxylation by P450s is depicted in Figure 8.
Chapter 1
General Introduction
Figure 8. The catalytic cycle of P450s, adapted from (66)
In the resting state, in absence of substrate, the ferric iron atom is six-coordinated,
complexed with a water molecule (I). Binding of a substrate displaces the water molecule
(II). This causes the ferric spin state to convert from low spin to high spin, a transition that
can be monitored spectrophotometrically since the Soret band undergoes a “type I” shift
from ca. 418 nm to ca. 390 nm. A single electron supplied by a reduced pyridine
nucleotide (NADPH or NADH) reduces the heme iron to the ferrous state (III).
Subsequently, dioxygen binds to the ferrous iron generating the oxy-complex (IV). A
second electron is then transferred, creating a ferric peroxy complex (V). Protonation of
the terminal oxygen atom gives “Compound 0”, a hydroxyperoxy adduct (VI). Addition of a
second hydrogen atom followed by loss of water gives “Compound I” (VII), a ferryl species
thought to be the active entity in most P450 oxidations, which has been characterized
recently (67). This complex abstracts hydrogen from the substrate (VIII) after which the
hydroxyl group formed rebounds to the substrate radical, resulting in formation of the
hydroxylated product which releases from the active site. NAD(P)H consumption is not
necessarily fully coupled to product formation (68). If dioxygen binding is hindered, the
second electron transfer will be too slow, resulting in the loss of superoxide from (IV)
(superoxide uncoupling). If substrate fitting in the active site cannot prevent water
encroachment, (VI) will be protonated at the iron-bound oxygen, resulting in the loss of
Chapter 1
General Introduction
hydrogen peroxide (peroxide uncoupling). If a substrate binds with no hydrogen atom
conveniently positioned for abstraction, the oxygen atom in (VII) could be reduced to
water (oxidase uncoupling). Those alternative pathways can hamper the use of P450s in
biocatalysis because the expensive electron donating cofactor is used in futile redox
cycling instead of product formation. Moreover the peroxide and superoxide formed by
uncoupling can damage the biocatalyst.
4.
Bacterial Cytochrome P450 BM3
Bacterial P450 systems often exhibit much higher catalytic activities than membranebound eukariotic P450s and are easy to handle in the laboratory due to their solubility and
high expression level in heterologous hosts (69). In particular Cytochrome P450 BM3
(BM3) is considered as one of the most promising monoxygenases for biotechnological
applications as it possesses the highest activity ever recorded for a P450 (63, 70, 71) and
can be easily expressed at high yield in Escherichia coli.
The natural substrates of BM3 are C12-C18 fatty acids which are hydroxylated at very high
activity at subterminal positions (72). However, over the recent years, several research
groups have succeeded in expanding the substrate selectivity of P450 BM3 and improving
its catalytic properties by site-directed and/or random mutagenesis. Through rational
redesign and directed evolution, BM3 mutants have been obtained that are able to oxidize
aromatics (73), alkanes (74), hydrocarbons (75), carboxylic acids (76) and pharmaceuticals
(21, 22, 77-84).
Recently a large body of research has been conducted to broaden and alter the substrate
specificity of bacterial CYPs in an effort to generate “human like” P450 activities (57). The
demand of “humanized” bacterial P450 with high activity is constantly growing to meet
the needs of the fine-chemical synthesis, pharmaceutical production and bioremediation
industries (85).
4.1. Engineering P450 BM3
The application of engineered bacterial P450s also have application in pharmacology and
toxicology as it enable the generation of substantial amounts of drug metabolites as well
as in optimization of lead compounds (86). Figure 9 shows examples for substrates of
engineered cytochrome P450 BM3 variants.
In order to improve the selectivity of oxidation to a specific product, the substrate from
which it is formed must be constrained to bind in a particular orientation (87). This can be
obtained by rational redesign of the active site, for example introducing polar residues or
bulky side-chains (88-90).
Chapter 1
General Introduction
Figure 9. Examples for substrates of engineered cytochrome P450 BM3 variants.
(Adapted from (62)).
Directed evolution, in which mutations are introduced at random positions using errorprone PCR, is a valuable complementary tool to site-directed mutagenesis. This approach
has the advantage of being able to generate unpredictable mutations with unanticipated
beneficial properties. The benefits of the two strategies can be combined using rational
evolution, in which selected residues are subjected to random or semi-random
mutagenesis. Alternatively, directed evolution can be used to identify sites suitable for
targeting, followed by site-saturation mutagenesis to establish which mutations are most
effective in those positions.
For example, Dietrich et al. (91) developed a novel semi biosynthetic route for the
production of artemisinin, a highly important antimalarian agent. By rational redesign of
Chapter 1
General Introduction
the active site they introduced mutations F87A/A328L in the active site of substrate
promiscuous P450 BM3: mutation F87A appeared to relieve the steric hindrance imposed
upon the substrate amorphadiene and allowed an increased access to the heme group,
while mutation A328L decreased the mobility of amorphadiene in the active site,
promoting epoxidation.
A similar approach has been used by Seifert et al. (75), who focused on the same hotspots
F87 and A328, creating a minimal and highly enriched P450 BM3 mutant library, by
mutating those two positions to all possible combinations of five hydrophobic aminoacids.
In this way, by systematically altering the size of the side chains in those two positions, a
broad range of binding site shapes was generated to convert a range of differently sized
and shaped terpenes (75). Recently, Seifert et al. (92) set up an iterative approach
comprising successive rounds of modeling, designing of focused mutant libraries and
screening, to identify a mutant for the production of the highly valuable product perillyl
alcohol.
It is widely accepted that increased regio-, stereoselectivity is the result of a reduced
number of substrate orientations close to the heme, as discussed in chapter 4 of the
present thesis (82, 90, 93).
Considerable effort has been invested for improving the coupling efficiency (ratio of
product formed to NAD(P)H cofactor consumed) of P450 BM3 by rational mutagenesis.
However the efficiencies obtained are still far lower than with native enzyme-substrate.
Detailed information on the structural determinants of coupling with novel substrates are
missing, therefore it is difficult to rationalize mutations to improve coupling efficiencies.
However, cost-effective, high-throughput screening methods will need to be developed to
assess relative coupling efficiency among P450s in mutant libraries (86).
Properties that need to be improved to enable the application of P450s in industrial
processes are: TTN (total turnover number), substrate acceptance, regio-and stereoselectivity, activity (kcat, KM), inhibition and coupling efficiency (70). Also improvement of
physical properties as thermostability, solvent tolerance, oxidative stability and substrate
and product tolerance should be addressed (94).
4.2. P450 BM3 in biocatalysis
One emerging application of P450s for the pharmaceutical industry is the generation of
large amount of drug metabolites, as discussed above (86). Several groups showed drug
metabolite preparation using human P450s expressed in Escherichia coli and in insect cells
(95). The Arnold group showed that human metabolites of propranolol could be produced
by mutants of bacterial P450 BM3 heme domain in a reaction driven by hydrogen
peroxide (5). Yields obtained were comparable to those produced by recombinant human
P450s in bacterial or baculovirus bioreactors (96). Recently, Dubey et al. set up a
biotechnological production of anticancer drug (colchicine derivatives) using P450 BM3 as
biocatalyst obtaining a yield of 7.3 g/L in 70 L fermentor (97). Kim at al. applied BM3 wt
Chapter 1
General Introduction
and a set of mutants for the generation of the human metabolite piceatannol from the
anticancer-preventive agent resveratrol (77), for the generation of human chiral
metabolites of simvastatin and lovastatin (79) and for the generation of human
metabolites of 7-ethoxycoumarin (78). The Arnold group compiled a panel of 120 mutants
to prepare nearly all human metabolites and a number of novel hydroxylated derivatives
for the drugs verapamil and astemizole (81). This panel of enzymes could also be applied
for the generation of new chemical entities (NCEs) to diversify lead compounds. Within
any catalyst panel both extremes of regioselectivity can be useful: enzymes selectively
producing individual metabolites and less selective enzymes to survey metabolite
possibilities (81).
A recent area of interest is the design and development of novel P450s as biocatalysts of
reactions of commercial interest, such as in pharmaceutical or fine chemical synthesis (86)
as well as for the optimization of lead compounds.
An example of the application of BM3 mutants in chemical synthesis is the
chemoenzymatic elaboration of monosaccharides applied by the Arnold group (98).
Regioselective deprotection of monosaccharide substrates using engineered P450 BM3
demethylases provides a highly efficient method to obtain valuable intermediates that can
be converted to a wide range of substituted monosaccharides and polysaccharides (98).
Engineered Cytochrome P450 BM3s were also applied for the enantioselective hydroxylation of 2-arylacetic acid derivatives and could produce an authentic human
metabolite of buspirone with high activity and selectivity (80).
Rentmeister et al. showed that products hydroxylated by BM3 can be subsequently
fluorinated in a chemo-enzymatic process that greatly simplifies the insertion of fluorine
into these structures at specific positions (99). Fluorination has become an important tool
in drug discovery and development as it can modulate the pharmacokinetic and the
pharmacological properties of drugs and lead compounds. In particular, fluorination can
improve membrane permeability, metabolic stability and/or receptor-binding properties
of bioactive molecules (99).
4.3. Key residues in the P450 BM3 active site
Control of regio- and stereoselectivity of biocatalysts is one of the major challenges in
biotechnology, as a limited number of residues in the substrate pocket appear to possess
significant selectivity-influencing powers. Ala82 and Phe87 are among those.
In Figure 10 key residues of the substrate channel and the active site of substrate-bound
P450 BM3 are shown. P450 BM3 crystallizes in a “precatalytic conformation” in which the
substrate is located too far to the heme iron for the oxidation to take place (87). Therefore
it is unclear whether the residues located in the substrate access channel are also
catalytically relevant. Also the spin relaxation NMR experiments of reduced P450 BM3
suggested that the substrates are 6 Å closer to the porphyrin ring than in the oxidized
form, indicating that structural rearrangements occur in order to allow the substrate to
Chapter 1
General Introduction
move closer to the heme iron (100). In the present thesis mutagenesis studies at position
87 (101) (84) and position 82 (89) will be presented.
Figure 10. The substrate access channel (A) and active site (B) of NPG-bound P450BM3
(Adapted from (66))
Phe 87. Phe 87 is one of the most studied residues of the active site of BM3 and it is object
of chapter 2 and 3 of the present thesis. In the substrate-free crystal structure, this
residue lies perpendicular to the porphyrin system, but in the substrate-bound crystal, the
Chapter 1
General Introduction
orientation of the aromatic ring changes to parallel to the plane of the heme (102) (103)
(63). It has been proposed that this motion influences the orientation of the substrate
relative to the catalytic center and it is responsible for the displacement of the axial water
ligand from its coordination site (104).
Phe87 forms with Phe81 a small hydrophobic pocket within the active site that sequesters
the -end of fatty acids. It seems likely that small hydrophobic substrates could
preferentially bind in the pocket generating non-productive complexes (105).
Phe87 is the most commonly mutated residue of P450 BM3 (Table 6).
Site-directed mutagenesis studies proved that Phe87 plays a key role in the control of
reaction selectivity (106) (107).
Typically, Phe87 was mutated to aminoacids with less bulky side chains, in order to
destroy the hydrophobic pocket, creating more space close to the heme, to accomodate
larger substrates (73-75, 77, 80, 103, 105, 108-116). Such substitutions, however, lead to a
decreased coupling efficiency of NADPH consumption to product formation due to a less
efficient exclusion of water from the active site (105).
The F87V mutation is a common component of several drug metabolizing variants (69)
(117). Moreover, it has been shown that mutation F87V significantly affected the
stereoselectivity of arachidonic acid epoxidation (106) and the activity towards indoles
(118). Li et al. also showed that mutation F87V is beneficial for oxidation of polycyclic
aromatic hydrocarbons, such as naphthalene, fluorene, acenaphthene, acenaphthylene,
and 9-methylanthracene (119).
In the present thesis saturation mutagenesis was applied to evaluate the effect of all
possible amino acids in this position on the metabolism of alkoxyresorufins, testosterone
and clozapine (84, 101).
Ala 82. The existence of a hydrophobic pocket between Phe81 and Phe87 clearly affects
the attempts to engineer BM3 towards novel target substrates. Huang et al. (90) were the
first to try to “fill” the hydrophobic pocket rather than destroying it, by mutating Ala82 to
large hydrophobic residues like Phe and Trp.
The mutants A82F and A82W showed a remarkable increase in affinity (800-fold) for fatty
acids. Moreover, the efficiency of indole hydroxylation increased, due both to increased
Kcat/KM values and to increased coupling efficiency, achieved by a more efficient exclusion
of water from the active site. Mutation at this position will be object of chapter 4 of the
present thesis.
Table 6. Reported mutations at position Phe87:
ALA (29 )(73-75) ( 81) (98, 99) (107 )(120-134)
ARG (101, 135)
ASN (101)
ASP (101)
Chapter 1
CYS
GLN
GLU
GLY
HIS
ILE
LEU
LYS
MET
PRO
SER
THR
TRP
TYR
VAL
SS
General Introduction
(101, 133)
(101, 135)
(101, 135)
(5, 74, 81, 98, 101, 122, 130, 135-143)
(101, 135)
(75, 81, 92, 98, 99, 135, 144, 145)
(75, 81, 92, 121, 124, 135, 144, 145)
(101, 135)
(101, 135)
(101, 135)
(101, 135)
(101, 135)
(101, 135)
(101) (106) (135) (136) (140) (146-147)
(69,74, 75, 81, 82, 98, 99, 106, 118, 119, 142, 144, 145,148-153)
(5, 91, 101, 118, 144, 145, 154, 155)
Table 7. Reported mutations at position Ala82:
CYS (81, 135, 144)
GLY (81, 99, 144, 155)
ILE
(81, 90, 144, 156)
LEU (5, 80, 81, 99, 121, 134, 144, 156, 157)
PHE (81, 90, 144, 156, 158)
PRO (5, 81)
SER (81, 144, 155)
THR (81, 144)
TRP (90, 99, 135, 159, 160)
VAL (156)
SS
(5, 144, 155)
5.
Spin lattice relaxation NMR (or T1 paramagnetic relaxation NMR)
In order to understand the structural basis that governs the regioselectivity of the
metabolism of a certain substrate or the specificity of a given P450 for a particular drug, it
1
is important to determine the structure of the substrate binding sites. For that purpose H
NMR appears as a powerful and attractive technique, since measurements of
Chapter 1
General Introduction
paramagnetic relaxation can be used to determine the distances between the heme iron
and protons of the bound substrate in P450-substrate complexes (161).
The heme iron atom of a CYP is paramagnetic when in oxidized state. Hydrogen atoms in a
magnetic field will re-align in the direction of this field after their orientation has been
flipped with a radio-frequency pulse. The velocity of this re-alignment (relaxation rate) is
dependent on the local strength of the magnetic field in the direct proximity of the
hydrogen atom. So a hydrogen atom close to the heme in the active site of a CYP, will
relax faster due to the magnetic moment of the heme iron atom.
The relationship between the iron atom to hydrogen atom distance and the rate of
relaxation has been established by Solomon and Bloembergen (162) and can be used to
measure substrate orientations in CYP active sites. A disadvantage of the technique is that
high concentrations (~mM) of ligands need to be present to measure NMR spectra.
Furthermore, average substrate active site orientations do not always match the
metabolic profile (163).
An overview of reported spin lattice relaxation studies using CYPs is given in Table 6.
In chapter 4 of the present thesis this technique has been successfully applied to
determine the orientation of testosterone in two BM3 mutants (M11 and M11 A82W) to
rationalize the different regioselectivity obtained in the metabolic profile.
Table 8. Spin lattice relaxation NMR studies used to determine active site ligand
orientations in CYPs.
CYP
102
Ligand
12-Br-lauric acid
1A
Paracetamol
1A2
Caffeine
2B
Paracetamol
Distances (A)*
7.8 (C1H)-16.3 (C12H)
5.9 (phenylH)6.7(methylH)
6.5 (N1methylH)6.7 (N3 and N7 methylH)
5.8 (methylH)- 6.3
(phenylH)
Major reaction
C2-OH
Benzoquinone
formation
Reference
(163)
N1-demethylation
(163)
Inactive
(164)
(164)
Chapter 1
General Introduction
*the minimal and maximal hydrogen atoms to heme iron atom measured are given.
**Measured in the F483I mutant of CYP2D6, the wild type is inactive towards
testosterone.
6.
Scope and objectives
The research presented in this thesis was part of a wider interdisciplinary project
“Metabolic stability assessment as a new tool in the Hit-to-Lead selection process and the
generation of new lead compound libraries” financed by the Dutch Top-Institute Pharma
consortium (grant D2-102). Several industrial (Merck/MSD (Oss, NL) and QPS
(Groeningen, NL)) and academic (Vrije Universiteit Amsterdam and Radboud Universiteit
Chapter 1
General Introduction
Nijmegen) partners participated in this project that led to several publications and four
PhD theses.
The aim of the project was the application of new, more efficient and selective P450 BM3
mutants for the generation of biotransformation products, being new chemical starting
points for lead optimization (lead libraries), followed by the determination of their binding
affinity (e.g. competitive fluorescence detection assay) and chemical structural
characterization (LC-MS/MS-NMR) in a hyphenated approach.
The general objectives of the present thesis are:
 Engineering drug metabolizing P450 BM3 mutants mimicking human P450s and
mutants with unique catalytic properties based on regio- and stereoselective
metabolism of diagnostic substrates.
 Utilization of “humanized” drug metabolizing P450 BM3 mutants for large scale
production of physiologically relevant human metabolites of lead compounds.
Figure 11 shows the general strategy applied in this research: by site-directed, random or
saturation mutagenesis BM3 mutants are engineered to obtain highly active mutants able
to metabolize drugs and drug-like molecules. The metabolic mixtures obtained by
incubation of diagnostic substrates with engineered BM3 mutants are then screened to
verify the acquisition of certain properties: HRS enables the selection of mutants able to
produce metabolites that show pharmacological activity, GSH trapping enables the
identification of mutants able to produce reactive metabolites (subsequently trapped by
GSH), screening by UPLC allows the identification of changes in regio- and stereoselectivity
in a quick and informative way.
Selected mutants that acquired a certain property (e.g. production of high amounts of a
pharmacologically active metabolite, high selectivity for the production of reactive
metabolites, high selectivity of the production of a certain hydroxy-steroid) are then
applied as biocatalysts for the large scale metabolite production in order to enable
pharmacological and toxicological evaluation of drug metabolites and for their structural
elucidation by NMR.
Chapter 1
General Introduction
Figure 11. General strategy applied for the research presented in this thesis, leading to a
“metabolite production and profiling platform”
7.
Outline of this thesis
The work presented in this thesis shows application of BM3 mutants as biocatalyst for
three purposes:
 For the regio- and stereo- selective hydroxylation of lead compounds (e.g.
steroids)
 For the generation of reactive metabolites
 For the generation of pharmacologically active metabolites / lead optimization
Regio- and stereo-selective hydroxylation of unactivated C-H bonds of natural or synthetic
compounds is one of the biggest challenges in synthetic organic chemistry. In particular
steroid hydroxylation is extremely challenging due to the complexity of the steroid
molecule that requires complicated, multi-step schemes involving protecting groups and
subsequent regeneration.
In chapter 2 and 4 of the present thesis cytochrome P450 BM3 mutants are
applied as biocatalysts for the regioselective hydroxylation of testosterone and
norethisterone. In chapter 2 a site saturation mutagenesis library of BM3 mutants was
applied to evaluate the effect of a single mutation in the active site of BM3 on the
regioselective hydroxylation of testosterone and on the coupling efficiency in
alkoxyresorufin oxidation. In chapter 4, a single active site substitution was applied in
order to reduce the substrate mobility of testosterone, thus improving the regioselectivity
Chapter 1
General Introduction
of hydroxylation. Spin lattice relaxation NMR was applied to determine how this single
mutation affected the orientation of the substrate within the active site of two BM3
mutants, M11 and M11 A82W.
The generation of reactive metabolites is important from a toxicological point of
view to be able to identify potentially toxic compounds that may be involved in ADRs or
IDRs.
Chapter 3 of the present thesis focuses on the application of P450 BM3 mutants
for the generation of reactive metabolites of clozapine, which is a drug known to be
responsible of severe ADRs. By screening of a minimal library of BM3 mutants, a mutant
was selected for the generation of large amounts of all major human relevant GSHconjugates of clozapine to enable their structural elucidation by NMR.
The generation of pharmacologically active metabolites is important from a “lead
optimization” point of view. Sometimes drug metabolites possess improved
pharmacological activity and improved physico-chemical properties, compared to the lead
itself.
In chapter 5 of the present thesis a focused library of BM3 mutants has been
applied for the generation of human relevant bioactive metabolites of p38 kinase
inhibitor Tak-715. The HRS screening applied enabled the identification and
pharmacological characterization of the drug metabolites in a quick and informative online
mode. Large scale incubation with BM3 mutants and isolation of active metabolites by
Prep-LC allowed their structural elucidation by NMR.
The last part of this thesis, chapter 6 concerns an overall summary of the work
described in the thesis including general conclusions and perspectives for future work.
References
Chapter 1
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
General Introduction
Fura, A., Shu, Y. Z., Zhu, M., Hanson, R. L., Roongta, V., and Humphreys, W. G.
(2004) Discovering drugs through biological transformation: role of
pharmacologically active metabolites in drug discovery, J Med Chem 47, 43394351.
Kang, M. J., Song, W. H., Shim, B. H., Oh, S. Y., Lee, H. Y., Chung, E. Y., Sohn, Y.,
and Lee, J. (2010) Pharmacologically active metabolites of currently marketed
drugs: potential resources for new drug discovery and development, Yakugaku
zasshi : Journal of the Pharmaceutical Society of Japan 130, 1325-1337.
Kumar, S. Engineering cytochrome P450 biocatalysts for biotechnology, medicine
and bioremediation, Expert Opin Drug Metab Toxicol 6, 115-131.
Schroer, K., Kittelmann, M., and Lutz, S. Recombinant human cytochrome P450
monooxygenases for drug metabolite synthesis, Biotechnol Bioeng 106, 699-706.
Otey, C. R., Bandara, G., Lalonde, J., Takahashi, K., and Arnold, F. H. (2006)
Preparation of human metabolites of propranolol using laboratory-evolved
bacterial cytochromes P450, Biotechnol Bioeng 93, 494-499.
www.fda.gov/downloads/Drugs/GuidanceComplianceRegulatoryInformation/
Guidances/ucm079266.pdf. (2008).
Nedderman, A. N. (2009) Metabolites in safety testing: metabolite identification
strategies in discovery and development, Biopharmaceutics & drug disposition 30,
153-162.
Evans, W. E., and Relling, M. V. (1999) Pharmacogenomics: translating functional
genomics into rational therapeutics, Science 286, 487-491.
Woolf, T. F., and Jordan, R. A. (1987) Basic concepts in drug metabolism: Part I,
Journal of clinical pharmacology 27, 15-17.
Jordan, R. A., and Woolf, T. F. (1987) Basic concepts in drug metabolism: Part II,
Journal of clinical pharmacology 27, 87-90.
Gomes, E. R., and Demoly, P. (2005) Epidemiology of hypersensitivity drug
reactions, Current opinion in allergy and clinical immunology 5, 309-316.
Williams, D. P., Kitteringham, N. R., Naisbitt, D. J., Pirmohamed, M., Smith, D.
A., and Park, B. K. (2002) Are chemically reactive metabolites responsible for
adverse reactions to drugs?, Current drug metabolism 3, 351-366.
Fontana, E., Dansette, P. M., and Poli, S. M. (2005) Cytochrome p450 enzymes
mechanism based inhibitors: common sub-structures and reactivity, Current drug
metabolism 6, 413-454.
Zhou, S., Chan, E., Lim, L. Y., Boelsterli, U. A., Li, S. C., Wang, J., Zhang, Q.,
Huang, M., and Xu, A. (2004) Therapeutic drugs that behave as mechanismbased inhibitors of cytochrome P450 3A4, Current drug metabolism 5, 415-442.
Caldwell, G. W., and Yan, Z. (2006) Screening for reactive intermediates and
toxicity assessment in drug discovery, Current opinion in drug discovery &
development 9, 47-60.
Prakash, C., Shaffer, C. L., and Nedderman, A. (2007) Analytical strategies for
identifying drug metabolites, Mass spectrometry reviews 26, 340-369.
DiMagno, E. P. (2007) The pancreatic cyst incidentaloma: management
consensus?, Clinical gastroenterology and hepatology : the official clinical
practice journal of the American Gastroenterological Association 5, 797-798.
Chapter 1
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
General Introduction
Madsen, K. G., Gronberg, G., Skonberg, C., Jurva, U., Hansen, S. H., and Olsen, J.
(2008) Electrochemical oxidation of troglitazone: identification and
characterization of the major reactive metabolite in liver microsomes, Chem Res
Toxicol 21, 2035-2041.
Chen, Y., Monshouwer, M., and Fitch, W. L. (2007) Analytical tools and
approaches for metabolite identification in early drug discovery, Pharm Res 24,
248-257.
Osorio-Lozada, A., Surapaneni, S., Skiles, G. L., and Subramanian, R. (2008)
Biosynthesis of drug metabolites using microbes in hollow fiber cartridge
reactors: case study of diclofenac metabolism by Actinoplanes species, Drug
Metab Dispos 36, 234-240.
Damsten, M. C., van Vugt-Lussenburg, B. M., Zeldenthuis, T., de Vlieger, J. S.,
Commandeur, J. N., and Vermeulen, N. P. (2008) Application of drug
metabolising mutants of cytochrome P450 BM3 (CYP102A1) as biocatalysts for
the generation of reactive metabolites, Chem Biol Interact 171, 96-107.
Damsten, M. C., de Vlieger, J. S., Niessen, W. M., Irth, H., Vermeulen, N. P., and
Commandeur, J. N. (2008) Trimethoprim: novel reactive intermediates and
bioactivation pathways by cytochrome p450s, Chem Res Toxicol 21, 2181-2187.
Boerma, J. S., Vermeulen, N. P., and Commandeur, J. N. (2011) Application of
CYP102A1M11H as a tool for the generation of protein adducts of reactive drug
metabolites, Chem Res Toxicol 24, 1263-1274.
Dragovic, S., Boerma, J. S., van Bergen, L., Vermeulen, N. P., and Commandeur,
J. N. (2010) Role of human glutathione S-transferases in the inactivation of
reactive metabolites of clozapine, Chem Res Toxicol 23, 1467-1476.
Lim, H. K., Stellingweif, S., Sisenwine, S., and Chan, K. W. (1999) Rapid drug
metabolite profiling using fast liquid chromatography, automated multiple-stage
mass spectrometry and receptor-binding, J Chromatogr A 831, 227-241.
Soldner, A., Spahn-Langguth, H., Palm, D., and Mutschler, E. (1998) A
radioreceptor assay for the analysis of AT1-receptor antagonists. Correlation with
complementary LC data reveals a potential contribution of active metabolites,
Journal of pharmaceutical and biomedical analysis 17, 111-124.
de Vlieger, J. S., Kolkman, A. J., Ampt, K. A., Commandeur, J. N., Vermeulen, N.
P., Kool, J., Wijmenga, S. S., Niessen, W. M., Irth, H., and Honing, M. (2010)
Determination and identification of estrogenic compounds generated with
biosynthetic enzymes using hyphenated screening assays, high resolution mass
spectrometry and off-line NMR, J Chromatogr B Analyt Technol Biomed Life Sci
878, 667-674.
Falck, D., de Vlieger, J. S., Niessen, W. M., Kool, J., Honing, M., Giera, M., and
Irth, H. (2010) Development of an online p38alpha mitogen-activated protein
kinase binding assay and integration of LC-HR-MS, Analytical and bioanalytical
chemistry 398, 1771-1780.
Reinen, J., Kalma, L. L., Begheijn, S., Heus, F., Commandeur, J. N., and
Vermeulen, N. P. (2011) Application of cytochrome P450 BM3 mutants as
biocatalysts for the profiling of estrogen receptor binding metabolites of the
mycotoxin zearalenone, Xenobiotica; the fate of foreign compounds in biological
systems 41, 59-70.
Chapter 1
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
General Introduction
Miller, K. J., and Herman, A. C. (1996) Affinity chromatography with
immunochemical detection applied to the analysis of human methionyl
granulocyte colony stimulating factor in serum, Anal Chem 68, 3077-3082.
Przyjazny, A., Hentz, N. G., and Bachas, L. G. (1993) Sensitive and selective liquid
chromatographic postcolumn reaction detection system for biotin and biocytin
using a homogeneous fluorophore-linked assay, J Chromatogr A 654, 79-86.
Oosterkamp, A. J., Irth, H., Beth, M., Unger, K. K., Tjaden, U. R., and van de
Greef, J. (1994) Bioanalysis of digoxin and its metabolites using direct serum
injection combined with liquid chromatography and on-line immunochemical
detection, Journal of chromatography. B, Biomedical applications 653, 55-61.
Kool, J., van Liempd, S. M., Ramautar, R., Schenk, T., Meerman, J. H., Irth, H.,
Commandeur, J. N., and Vermeulen, N. P. (2005) Development of a novel
cytochrome p450 bioaffinity detection system coupled online to gradient
reversed-phase high-performance liquid chromatography, Journal of
biomolecular screening 10, 427-436.
Ingkaninan, K., Hazekamp, A., de Best, C. M., Irth, H., Tjaden, U. R., van der
Heijden, R., van der Greef, J., and Verpoorte, R. (2000) The application of HPLC
with on-line coupled UV/MS-biochemical detection for isolation of an
acetylcholinesterase inhibitor from narcissus 'Sir Winston Churchill', J Nat Prod
63, 803-806.
Schenk, T., Appels, N. M., van Elswijk, D. A., Irth, H., Tjaden, U. R., and van der
Greef, J. (2003) A generic assay for phosphate-consuming or -releasing enzymes
coupled on-line to liquid chromatography for lead finding in natural products,
Anal Biochem 316, 118-126.
Krabbe, J. G., Lingeman, H., Niessen, W. M., and Irth, H. (2003) Ligand-exchange
detection of phosphorylated peptides using liquid chromatography electrospray
mass spectrometry, Anal Chem 75, 6853-6860.
Van Liempd, S. M., Kool, J., Meerman, J. H., Irth, H., and Vermeulen, N. P. (2007)
Metabolic profiling of endocrine-disrupting compounds by on-line cytochrome
p450 bioreaction coupled to on-line receptor affinity screening, Chem Res Toxicol
20, 1825-1832.
Giera, M., de Vlieger, J. S., Lingeman, H., Irth, H., and Niessen, W. M. (2010)
Structural elucidation of biologically active neomycin N-octyl derivatives in a
regioisomeric mixture by means of liquid chromatography/ion trap time-of-flight
mass spectrometry, Rapid communications in mass spectrometry : RCM 24, 14391446.
Giera, M., Heus, F., Janssen, L., Kool, J., Lingeman, H., and Irth, H. (2009)
Microfractionation revisited: a 1536 well high resolution screening assay, Anal
Chem 81, 5460-5466.
Donova, M. V., and Egorova, O. V. (2012) Microbial steroid transformations:
current state and prospects, Appl Microbiol Biotechnol 94, 1423-1447.
Hogg, J. A. (1992) Steroids, the steroid community, and Upjohn in perspective: a
profile of innovation, Steroids 57, 593-616.
Sedlaczek, L. (1988) Biotransformations of steroids, Crit Rev Biotechnol 7, 187236.
Chapter 1
43.
44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
General Introduction
Fernandes, A., Cruz, A., Angelova, B., Pinheiro, H. M., and Cabral, A. (2003)
Microbial conversion of steroid compounds: recent developments, Enzyme
Microb Technol 32, 688-705.
Mahato, S. B., Banerjee, S., and Sahu, N. P. (1984) Metabolism of progesterone
and testosterone by a Bacillus sp, Steroids 43, 545-558.
Mahato, S. B., and Mukherjee, A. (1984) Microbial transformation of
testosterone by Aspergillus fumigatus, J Steroid Biochem 21, 341-342.
Ko, D., Heiman, A. S., Chen, M., and Lee, H. J. (2000) New steroidal antiinflammatory antedrugs: methyl 21-desoxy-21-chloro-11beta,17alpha-dihydroxy3,20-dioxo-1, 4-pregnadiene-16alpha-carboxylate, methyl 21-desoxy-21-chloro11beta-hydroxy-3,20-dioxo-1, 4-pregnadiene-16alpha-carboxylate, and their
9alpha-fluoro derivatives*, Steroids 65, 210-218.
Tuba, Z., Bardin, C. W., Dancsi, A., Francsics-Czinege, E., Molnar, C., Csorgei, J.,
Falkay, G., Koide, S. S., Kumar, N., Sundaram, K., Dukat-Abrok, V., and Balogh,
G. (2000) Synthesis and biological activity of a new progestogen, 16-methylene17alpha-hydroxy-18-methyl-19-norpregn-4-ene-3, 20-dione acetate, Steroids 65,
266-274.
Newhouse, T., and Baran, P. S. (2011) If C-H bonds could talk: selective C-H bond
oxidation, Angew Chem Int Ed Engl 50, 3362-3374.
Dutton, D. R., McMillen, S. K., Sonderfan, A. J., Thomas, P. E., and Parkinson, A.
(1987) Studies on the rate-determining factor in testosterone hydroxylation by
rat liver microsomal cytochrome P-450: evidence against cytochrome P-450
isozyme:isozyme interactions, Archives of biochemistry and biophysics 255, 316328.
Sonderfan, A. J., Arlotto, M. P., and Parkinson, A. (1989) Identification of the
cytochrome P-450 isozymes responsible for testosterone oxidation in rat lung,
kidney, and testis: evidence that cytochrome P-450a (P450IIA1) is the
physiologically important testosterone 7 alpha-hydroxylase in rat testis,
Endocrinology 125, 857-866.
Agematu, H., Matsumoto, N., Fujii, Y., Kabumoto, H., Doi, S., Machida, K.,
Ishikawa, J., and Arisawa, A. (2006) Hydroxylation of testosterone by bacterial
cytochromes P450 using the Escherichia coli expression system, Bioscience,
biotechnology, and biochemistry 70, 307-311.
Bernhardt, R. (2006) Cytochromes P450 as versatile biocatalysts, J Biotechnol
124, 128-145.
Hannemann, F., Bichet, A., Ewen, K. M., and Bernhardt, R. (2007) Cytochrome
P450 systems--biological variations of electron transport chains, Biochim Biophys
Acta 1770, 330-344.
Novikova, L. A., Faletrov, Y. V., Kovaleva, I. E., Mauersberger, S., Luzikov, V. N.,
and Shkumatov, V. M. (2009) From structure and functions of steroidogenic
enzymes to new technologies of gene engineering, Biochemistry. Biokhimiia 74,
1482-1504.
Kristan, K., and Rizner, T. L. (2012) Steroid-transforming enzymes in fungi, The
Journal of steroid biochemistry and molecular biology 129, 79-91.
Guengerich, F. P. (2001) Common and uncommon cytochrome P450 reactions
related to metabolism and chemical toxicity, Chem Res Toxicol 14, 611-650.
Chapter 1
57.
58.
59.
60.
61.
62.
63.
64.
65.
66.
67.
68.
69.
70.
71.
72.
73.
General Introduction
O'Reilly, E., Kohler, V., Flitsch, S. L., and Turner, N. J. (2010) Cytochromes P450
as useful biocatalysts: addressing the limitations, Chem Commun (Camb) 47,
2490-2501.
Drnelson.utmem.edu.
Pirmohamed, M., and Park, B. K. (2003) Cytochrome P450 enzyme
polymorphisms and adverse drug reactions, Toxicology 192, 23-32.
Ingelman-Sundberg, M. (2004) Pharmacogenetics of cytochrome P450 and its
applications in drug therapy: the past, present and future, Trends in
pharmacological sciences 25, 193-200.
Kirchheiner, J., and Brockmoller, J. (2005) Clinical consequences of cytochrome
P450 2C9 polymorphisms, Clin Pharmacol Ther 77, 1-16.
Urlacher, V. B., and Girhard, M. (2011) Cytochrome P450 monooxygenases: an
update on perspectives for synthetic application, Trends Biotechnol 30, 26-36.
Munro, A. W., Leys, D. G., McLean, K. J., Marshall, K. R., Ost, T. W., Daff, S.,
Miles, C. S., Chapman, S. K., Lysek, D. A., Moser, C. C., Page, C. C., and Dutton, P.
L. (2002) P450 BM3: the very model of a modern flavocytochrome, Trends
Biochem Sci 27, 250-257.
Isin, E. M., and Guengerich, F. P. (2007) Complex reactions catalyzed by
cytochrome P450 enzymes, Biochim Biophys Acta 1770, 314-329.
Anzenbacher, P., and Anzenbacherova, E. (2001) Cytochromes P450 and
metabolism of xenobiotics, Cell Mol Life Sci 58, 737-747.
Whitehouse, C. J., Bell, S. G., and Wong, L. L. (2012) P450(BM3) (CYP102A1):
connecting the dots, Chemical Society reviews 41, 1218-1260.
Rittle, J., and Green, M. T. (2010) Cytochrome P450 compound I: capture,
characterization, and C-H bond activation kinetics, Science 330, 933-937.
Loida, P. J., and Sligar, S. G. (1993) Molecular recognition in cytochrome P-450:
mechanism for the control of uncoupling reactions, Biochemistry 32, 1153011538.
Lussenburg, B. M., Babel, L. C., Vermeulen, N. P., and Commandeur, J. N. (2005)
Evaluation of alkoxyresorufins as fluorescent substrates for cytochrome P450
BM3 and site-directed mutants, Anal Biochem 341, 148-155.
Jung, S. T., Lauchli, R., and Arnold, F. H. (2011) Cytochrome P450: taming a wild
type enzyme, Curr Opin Biotechnol 22, 809-817.
Warman, A. J., Roitel, O., Neeli, R., Girvan, H. M., Seward, H. E., Murray, S. A.,
McLean, K. J., Joyce, M. G., Toogood, H., Holt, R. A., Leys, D., Scrutton, N. S., and
Munro, A. W. (2005) Flavocytochrome P450 BM3: an update on structure and
mechanism of a biotechnologically important enzyme, Biochem Soc Trans 33,
747-753.
Narhi, L. O., and Fulco, A. J. (1986) Characterization of a catalytically selfsufficient 119,000-dalton cytochrome P-450 monooxygenase induced by
barbiturates in Bacillus megaterium, J Biol Chem 261, 7160-7169.
Carmichael, A. B., and Wong, L. L. (2001) Protein engineering of Bacillus
megaterium CYP102. The oxidation of polycyclic aromatic hydrocarbons, Eur J
Biochem 268, 3117-3125.
Chapter 1
74.
75.
76.
77.
78.
79.
80.
81.
82.
83.
84.
85.
86.
General Introduction
Dietrich, M., Do, T. A., Schmid, R. D., Pleiss, J., and Urlacher, V. B. (2009)
Altering the regioselectivity of the subterminal fatty acid hydroxylase P450 BM-3
towards gamma- and delta-positions, J Biotechnol 139, 115-117.
Seifert, A., Vomund, S., Grohmann, K., Kriening, S., Urlacher, V. B., Laschat, S.,
and Pleiss, J. (2009) Rational design of a minimal and highly enriched CYP102A1
mutant library with improved regio-, stereo- and chemoselectivity, Chembiochem
10, 853-861.
Munzer, D. F., Meinhold, P., Peters, M. W., Feichtenhofer, S., Griengl, H.,
Arnold, F. H., Glieder, A., and de Raadt, A. (2005) Stereoselective hydroxylation
of an achiral cyclopentanecarboxylic acid derivative using engineered P450s BM3, Chem Commun (Camb), 2597-2599.
Kim, D. H., Ahn, T., Jung, H. C., Pan, J. G., and Yun, C. H. (2009) Generation of the
human metabolite piceatannol from the anticancer-preventive agent resveratrol
by bacterial cytochrome P450 BM3, Drug Metab Dispos 37, 932-936.
Kim, D. H., Kim, K. H., Liu, K. H., Jung, H. C., Pan, J. G., and Yun, C. H. (2008)
Generation of human metabolites of 7-ethoxycoumarin by bacterial cytochrome
P450 BM3, Drug Metab Dispos 36, 2166-2170.
Kim, K. H., Kang, J. Y., Kim, D. H., Park, S. H., Kim, D., Park, K. D., Lee, Y. J., Jung,
H. C., Pan, J. G., Ahn, T., and Yun, C. H. (2011) Generation of human chiral
metabolites of simvastatin and lovastatin by bacterial CYP102A1 mutants, Drug
Metab Dispos 39, 140-150.
Landwehr, M., Hochrein, L., Otey, C. R., Kasrayan, A., Backvall, J. E., and Arnold,
F. H. (2006) Enantioselective alpha-hydroxylation of 2-arylacetic acid derivatives
and buspirone catalyzed by engineered cytochrome P450 BM-3, J Am Chem Soc
128, 6058-6059.
Sawayama, A. M., Chen, M. M., Kulanthaivel, P., Kuo, M. S., Hemmerle, H., and
Arnold, F. H. (2009) A panel of cytochrome P450 BM3 variants to produce drug
metabolites and diversify lead compounds, Chemistry 15, 11723-11729.
van Vugt-Lussenburg, B. M., Damsten, M. C., Maasdijk, D. M., Vermeulen, N. P.,
and Commandeur, J. N. (2006) Heterotropic and homotropic cooperativity by a
drug-metabolising mutant of cytochrome P450 BM3, Biochem Biophys Res
Commun 346, 810-818.
van Vugt-Lussenburg, B. M., Stjernschantz, E., Lastdrager, J., Oostenbrink, C.,
Vermeulen, N. P., and Commandeur, J. N. (2007) Identification of critical
residues in novel drug metabolizing mutants of cytochrome P450 BM3 using
random mutagenesis, J Med Chem 50, 455-461.
Rea, V., Dragovic, S., Boerma, J. S., de Kanter, F. J., Vermeulen, N. P., and
Commandeur, J. N. (2011) Role of residue 87 in the activity and regioselectivity of
clozapine metabolism by drug-metabolizing CYP102A1 M11H: application for
structural characterization of clozapine GSH conjugates, Drug Metab Dispos 39,
2411-2420.
Yun, C. H., Kim, K. H., Kim, D. H., Jung, H. C., and Pan, J. G. (2007) The bacterial
P450 BM3: a prototype for a biocatalyst with human P450 activities, Trends
Biotechnol 25, 289-298.
Gillam, E. M. (2008) Engineering cytochrome p450 enzymes, Chem Res Toxicol
21, 220-231.
Chapter 1
87.
88.
89.
90.
91.
92.
93.
94.
95.
96.
97.
98.
General Introduction
Papouchado, B. G., Myles, J., Lloyd, R. V., Stoler, M., Oliveira, A. M., DownsKelly, E., Morey, A., Bilous, M., Nagle, R., Prescott, N., Wang, L., Dragovich, L.,
McElhinny, A., Garcia, C. F., Ranger-Moore, J., Free, H., Powell, W., Loftus, M.,
Pettay, J., Gaire, F., Roberts, C., Dietel, M., Roche, P., Grogan, T., and Tubbs, R.
(2010) Silver in situ hybridization (SISH) for determination of HER2 gene status in
breast carcinoma: comparison with FISH and assessment of interobserver
reproducibility, The American journal of surgical pathology 34, 767-776.
Feenstra, K. A., Hofstetter, K., Bosch, R., Schmid, A., Commandeur, J. N., and
Vermeulen, N. P. (2006) Enantioselective substrate binding in a monooxygenase
protein model by molecular dynamics and docking, Biophysical journal 91, 32063216.
Rea, V., Kolkman, A. J., Vottero, E. R., Stronks, E. J., Ampt, K. A., Honing, M.,
Vermeulen, N. P., Wijmenga, S. S., and Commandeur, J. N. (2011) Active site
substitution A82W improves regioselectivity of steroid hydroxylation by
cytochrome P450 BM3 mutants as rationalised by spin relaxation NMR studies,
Biochemistry.
Huang, W. C., Westlake, A. C., Marechal, J. D., Joyce, M. G., Moody, P. C., and
Roberts, G. C. (2007) Filling a hole in cytochrome P450 BM3 improves substrate
binding and catalytic efficiency, J Mol Biol 373, 633-651.
Dietrich, J. A., Yoshikuni, Y., Fisher, K. J., Woolard, F. X., Ockey, D., McPhee, D.
J., Renninger, N. S., Chang, M. C., Baker, D., and Keasling, J. D. (2009) A novel
semi-biosynthetic route for artemisinin production using engineered substratepromiscuous P450(BM3), ACS chemical biology 4, 261-267.
Seifert, A., Antonovici, M., Hauer, B., and Pleiss, J. (2011) An efficient route to
selective bio-oxidation catalysts: an iterative approach comprising modeling,
diversification, and screening, based on CYP102A1, Chembiochem 12, 1346-1351.
Koenig, R., Lesemann, D. E., Loss, S., Engelmann, J., Commandeur, U., Deml, G.,
Schiemann, J., Aust, H., and Burgermeister, W. (2006) Zygocactus virus X-based
expression vectors and formation of rod-shaped virus-like particles in plants by
the expressed coat proteins of Beet necrotic yellow vein virus and Soil-borne
cereal mosaic virus, The Journal of general virology 87, 439-443.
Julsing, M. K., Cornelissen, S., Buhler, B., and Schmid, A. (2008) Heme-iron
oxygenases: powerful industrial biocatalysts?, Curr Opin Chem Biol 12, 177-186.
Parikh, A., Gillam, E. M., and Guengerich, F. P. (1997) Drug metabolism by
Escherichia coli expressing human cytochromes P450, Nat Biotechnol 15, 784788.
Rushmore, T. H., Reider, P. J., Slaughter, D., Assang, C., and Shou, M. (2000)
Bioreactor systems in drug metabolism: synthesis of cytochrome P450-generated
metabolites, Metab Eng 2, 115-125.
Dubey, K. K., and Behera, B. K. (2011) Statistical optimization of process variables
for the production of an anticancer drug (colchicine derivatives) through
fermentation: at scale-up level, New biotechnology 28, 79-85.
Lewis, J. C., Bastian, S., Bennett, C. S., Fu, Y., Mitsuda, Y., Chen, M. M.,
Greenberg, W. A., Wong, C. H., and Arnold, F. H. (2009) Chemoenzymatic
elaboration of monosaccharides using engineered cytochrome P450BM3
demethylases, Proc Natl Acad Sci U S A 106, 16550-16555.
Chapter 1
99.
100.
101.
102.
103.
104.
105.
106.
107.
108.
109.
110.
111.
General Introduction
Rentmeister, A., Arnold, F. H., and Fasan, R. (2009) Chemo-enzymatic
fluorination of unactivated organic compounds, Nat Chem Biol 5, 26-28.
Ferrero, V. E., Di Nardo, G., Catucci, G., Sadeghi, S. J., and Gilardi, G. (2012)
Fluorescence detection of ligand binding to labeled cytochrome P450 BM3,
Dalton Trans 41, 2018-2025.
Vottero, E., Rea, V., Lastdrager, J., Honing, M., Vermeulen, N. P., and
Commandeur, J. N. (2011) Role of residue 87 in substrate selectivity and
regioselectivity of drug-metabolizing cytochrome P450 CYP102A1 M11, J Biol
Inorg Chem 16, 899-912.
Okubo, Y., Bessho, K., Fujimura, K., Kusumoto, K., Ogawa, Y., and Iizuka, T.
(2002) Expression of bone morphogenetic protein in the course of osteoinduction
by recombinant human bone morphogenetic protein-2, Clinical oral implants
research 13, 80-85.
Ogawa, M., Kumagai, K., Gondo, N., Matsumoto, N., Suyama, K., and Saku, K.
(2002) Novel electrophysiologic parameter of dispersion of atrial repolarization:
comparison of different atrial pacing methods, J Cardiovasc Electrophysiol 13,
110-117.
Fedorova, A., and Ogawa, M. Y. (2002) Site-specific modification of de novo
designed coiled-coil polypeptides with inorganic redox complexes, Bioconjug
Chem 13, 150-154.
Hoehn, R., Lamparter, J., Pfeiffer, N., and Vossmerbaeumer, U. (2011) Seizures
following subconjunctival 5-FU therapy, Graefe's archive for clinical and
experimental ophthalmology = Albrecht von Graefes Archiv fur klinische und
experimentelle Ophthalmologie 249, 145-146.
Graham-Lorence, S., Truan, G., Peterson, J. A., Falck, J. R., Wei, S., Helvig, C.,
and Capdevila, J. H. (1997) An active site substitution, F87V, converts cytochrome
P450 BM-3 into a regio- and stereoselective (14S,15R)-arachidonic acid
epoxygenase, J Biol Chem 272, 1127-1135.
Oliver, C. F., Modi, S., Sutcliffe, M. J., Primrose, W. U., Lian, L. Y., and Roberts,
G. C. (1997) A single mutation in cytochrome P450 BM3 changes substrate
orientation in a catalytic intermediate and the regiospecificity of hydroxylation,
Biochemistry 36, 1567-1572.
Nulens, E., Gould, I., MacKenzie, F., Deplano, A., Cookson, B., Alp, E., Bouza, E.,
and Voss, A. (2005) Staphylococcus aureus carriage among participants at the
13th European Congress of Clinical Microbiology and Infectious Diseases,
European journal of clinical microbiology & infectious diseases : official
publication of the European Society of Clinical Microbiology 24, 145-148.
Schotten, U., Schumacher, C., Conrads, V., Braun, V., Schondube, F., Voss, M.,
and Hanrath, P. (1999) Calcium-sensitivity of the SR calcium release channel in
failing and nonfailing human myocardium, Basic research in cardiology 94, 145151.
de Smet, M. D., Buffam, F. V., Fairey, R. N., and Voss, N. J. (1990) Prevention of
radiation-induced stenosis of the nasolacrimal duct, Canadian journal of
ophthalmology. Journal canadien d'ophtalmologie 25, 145-147.
Voss, K., Barz, H., and Wagner, H. (1977) [Image processing in pathology. III.
Structure and use of an automatic morphometry system for recording
Chapter 1
112.
113.
114.
115.
116.
117.
118.
119.
120.
121.
122.
123.
124.
General Introduction
morphometric features of nuclei (author's transl)], Experimentelle Pathologie 13,
145-161.
Voss, R., Rickart, A., Ruile, K., Schoen, H. R., and Kunz, W. (1970) [Hyperbaric
organ preservation], Verhandlungen der Deutschen Gesellschaft fur Pathologie
54, 145-150.
Vossschulte, K. (1951) [The intervertebral disks as affected by the movements of
the spine], Langenbecks Archiv fur klinische Chirurgie ... vereinigt mit Deutsche
Zeitschrift fur Chirurgie 267, 145.
Ruperti, B., Bonghi, C., Rasori, A., Ramina, A., and Tonutti, P. (2001)
Characterization and expression of two members of the peach 1aminocyclopropane-1-carboxylate oxidase gene family, Physiol Plant 111, 336344.
Bidoli, V., Casolino, M., De Pascale, M. P., Furano, G., Minori, M., Morselli, A.,
Narici, L., Picozza, P., Reali, E., Sparvoli, R., Fuglesang, C., Sannita, W., Carlson,
P., Castellini, G., Galper, A., Korotkov, M., Popov, A., Navilov, N., Avdeev, S.,
Benghin, V., Salnitskii, V., Shevchenko, O., Boezio, M., Bonvicini, W., Vacchi, A.,
Zampa, G., Zampa, N., Mazzenga, G., Ricci, M., Spillantini, P., and Vittori, R.
(2002) The Sileye-3/Alteino experiment for the study of light flashes, radiation
environment and astronaut brain activity on board the International Space
Station, Journal of radiation research 43 Suppl, S47-52.
Furalev, V. A., Kravchenko, I. V., Vasil'eva, L. A., and Pylev, L. N. (2002) Cytotoxic
effects of macrophages and asbestos on transformed rat mesothelial cells,
Bulletin of experimental biology and medicine 133, 71-73.
Walldal, E., Anund, I., and Furaker, C. (2002) Quality of care and development of
a critical pathway, Journal of nursing management 10, 115-122.
Li, Q. S., Schwaneberg, U., Fischer, P., and Schmid, R. D. (2000) Directed
evolution of the fatty-acid hydroxylase P450 BM-3 into an indole-hydroxylating
catalyst, Chemistry 6, 1531-1536.
Li, Q. S., Ogawa, J., Schmid, R. D., and Shimizu, S. (2001) Engineering cytochrome
P450 BM-3 for oxidation of polycyclic aromatic hydrocarbons, Appl Environ
Microbiol 67, 5735-5739.
Volz, T. J., Rock, D. A., and Jones, J. P. (2002) Evidence for two different active
oxygen species in cytochrome P450 BM3 mediated sulfoxidation and Ndealkylation reactions, J Am Chem Soc 124, 9724-9725.
Chen, C. K., Shokhireva, T., Berry, R. E., Zhang, H., and Walker, F. A. (2008) The
effect of mutation of F87 on the properties of CYP102A1-CYP4C7 chimeras:
altered regiospecificity and substrate selectivity, J Biol Inorg Chem 13, 813-824.
Eiben, S., Kaysser, L., Maurer, S., Kuhnel, K., Urlacher, V. B., and Schmid, R. D.
(2006) Preparative use of isolated CYP102 monooxygenases -- a critical appraisal,
J Biotechnol 124, 662-669.
Rock, D. A., Perkins, B. N., Wahlstrom, J., and Jones, J. P. (2003) A method for
determining two substrates binding in the same active site of cytochrome
P450BM3: an explanation of high energy omega product formation, Archives of
biochemistry and biophysics 416, 9-16.
Whitehouse, C. J., Yang, W., Yorke, J. A., Tufton, H. G., Ogilvie, L. C., Bell, S. G.,
Zhou, W., Bartlam, M., Rao, Z., and Wong, L. L. (2011) Structure, electronic
Chapter 1
125.
126.
127.
128.
129.
130.
131.
132.
133.
134.
135.
136.
137.
General Introduction
properties and catalytic behaviour of an activity-enhancing CYP102A1
(P450(BM3)) variant, Dalton Trans 40, 10383-10396.
Haines, D. C., Tomchick, D. R., Machius, M., and Peterson, J. A. (2001) Pivotal
role of water in the mechanism of P450BM-3, Biochemistry 40, 13456-13465.
Rock, D. A., Boitano, A. E., Wahlstrom, J. L., and Jones, J. P. (2002) Use of kinetic
isotope effects to delineate the role of phenylalanine 87 in P450(BM-3),
Bioorganic chemistry 30, 107-118.
Feenstra, K. A., Starikov, E. B., Urlacher, V. B., Commandeur, J. N., and
Vermeulen, N. P. (2007) Combining substrate dynamics, binding statistics, and
energy barriers to rationalize regioselective hydroxylation of octane and lauric
acid by CYP102A1 and mutants, Protein science : a publication of the Protein
Society 16, 420-431.
Wong, T. S., Arnold, F. H., and Schwaneberg, U. (2004) Laboratory evolution of
cytochrome p450 BM-3 monooxygenase for organic cosolvents, Biotechnol
Bioeng 85, 351-358.
Roccatano, D., Wong, T. S., Schwaneberg, U., and Zacharias, M. (2006) Toward
understanding the inactivation mechanism of monooxygenase P450 BM-3 by
organic cosolvents: a molecular dynamics simulation study, Biopolymers 83, 467476.
Urlacher, V. B., Makhsumkhanov, A., and Schmid, R. D. (2006)
Biotransformation of beta-ionone by engineered cytochrome P450 BM-3, Appl
Microbiol Biotechnol 70, 53-59.
Sowden, R. J., Yasmin, S., Rees, N. H., Bell, S. G., and Wong, L. L. (2005)
Biotransformation of the sesquiterpene (+)-valencene by cytochrome P450cam
and P450BM-3, Org Biomol Chem 3, 57-64.
Whitehouse, C. J., Bell, S. G., Tufton, H. G., Kenny, R. J., Ogilvie, L. C., and Wong,
L. L. (2008) Evolved CYP102A1 (P450BM3) variants oxidise a range of non-natural
substrates and offer new selectivity options, Chem Commun (Camb), 966-968.
Wong, T. S., Wu, N., Roccatano, D., Zacharias, M., and Schwaneberg, U. (2005)
Sensitive assay for laboratory evolution of hydroxylases toward aromatic and
heterocyclic compounds, Journal of biomolecular screening 10, 246-252.
Chen, C. K., Berry, R. E., Shokhireva, T., Murataliev, M. B., Zhang, H., and
Walker, F. A. (2010) Scanning chimeragenesis: the approach used to change the
substrate selectivity of fatty acid monooxygenase CYP102A1 to that of terpene
omega-hydroxylase CYP4C7, J Biol Inorg Chem 15, 159-174.
Reinen, J., Ferman, S., Vottero, E., Vermeulen, N. P., and Commandeur, J. N.
(2011) Application of a fluorescence-based continuous-flow bioassay to screen
for diversity of cytochrome P450 BM3 mutant libraries, Journal of biomolecular
screening 16, 239-250.
Noble, M. A., Miles, C. S., Chapman, S. K., Lysek, D. A., MacKay, A. C., Reid, G.
A., Hanzlik, R. P., and Munro, A. W. (1999) Roles of key active-site residues in
flavocytochrome P450 BM3, Biochem J 339 ( Pt 2), 371-379.
Raner, G. M., Hatchell, J. A., Dixon, M. U., Joy, T. L., Haddy, A. E., and Johnston,
E. R. (2002) Regioselective peroxo-dependent heme alkylation in P450(BM3)F87G by aromatic aldehydes: effects of alkylation on cataysis, Biochemistry 41,
9601-9610.
Chapter 1
138.
139.
140.
141.
142.
143.
144.
145.
146.
147.
148.
149.
150.
General Introduction
Brenner, S., Hay, S., Girvan, H. M., Munro, A. W., and Scrutton, N. S. (2007)
Conformational dynamics of the cytochrome P450 BM3/N-palmitoylglycine
complex: the proposed "proximal-distal" transition probed by temperature-jump
spectroscopy, J Phys Chem B 111, 7879-7886.
Raner, G. M., Thompson, J. I., Haddy, A., Tangham, V., Bynum, N., Ramachandra
Reddy, G., Ballou, D. P., and Dawson, J. H. (2006) Spectroscopic investigations of
intermediates in the reaction of cytochrome P450(BM3)-F87G with surrogate
oxygen atom donors, J Inorg Biochem 100, 2045-2053.
Noble, M. A., Quaroni, L., Chumanov, G. D., Turner, K. L., Chapman, S. K.,
Hanzlik, R. P., and Munro, A. W. (1998) Imidazolyl carboxylic acids as mechanistic
probes of flavocytochrome P-450 BM3, Biochemistry 37, 15799-15807.
Branco, R. J., Seifert, A., Budde, M., Urlacher, V. B., Ramos, M. J., and Pleiss, J.
(2008) Anchoring effects in a wide binding pocket: the molecular basis of
regioselectivity in engineered cytochrome P450 monooxygenase from B.
megaterium, Proteins 73, 597-607.
Li, Q. S., Ogawa, J., Schmid, R. D., and Shimizu, S. (2001) Residue size at position
87 of cytochrome P450 BM-3 determines its stereoselectivity in propylbenzene
and 3-chlorostyrene oxidation, FEBS Lett 508, 249-252.
Waltham, T. N., Girvan, H. M., Butler, C. F., Rigby, S. R., Dunford, A. J., Holt, R.
A., and Munro, A. W. (2011) Analysis of the oxidation of short chain alkynes by
flavocytochrome P450 BM3, Metallomics 3, 369-378.
Meinhold, P., Peters, M. W., Chen, M. M., Takahashi, K., and Arnold, F. H.
(2005) Direct conversion of ethane to ethanol by engineered cytochrome P450
BM3, Chembiochem 6, 1765-1768.
Kubo, T., Peters, M. W., Meinhold, P., and Arnold, F. H. (2006) Enantioselective
epoxidation of terminal alkenes to (R)- and (S)-epoxides by engineered
cytochromes P450 BM-3, Chemistry 12, 1216-1220.
Kitazume, T., Haines, D. C., Estabrook, R. W., Chen, B., and Peterson, J. A. (2007)
Obligatory intermolecular electron-transfer from FAD to FMN in dimeric P450BM3, Biochemistry 46, 11892-11901.
Daiber, A., Herold, S., Schoneich, C., Namgaladze, D., Peterson, J. A., and Ullrich,
V. (2000) Nitration and inactivation of cytochrome P450BM-3 by peroxynitrite.
Stopped-flow measurements prove ferryl intermediates, Eur J Biochem 267,
6729-6739.
Cowart, L. A., Falck, J. R., and Capdevila, J. H. (2001) Structural determinants of
active site binding affinity and metabolism by cytochrome P450 BM-3, Archives of
biochemistry and biophysics 387, 117-124.
Li, Q. S., Ogawa, J., Schmid, R. D., and Shimizu, S. (2005) Indole hydroxylation by
bacterial cytochrome P450 BM-3 and modulation of activity by cumene
hydroperoxide, Bioscience, biotechnology, and biochemistry 69, 293-300.
Lewis, J. C., Mantovani, S. M., Fu, Y., Snow, C. D., Komor, R. S., Wong, C. H., and
Arnold, F. H. (2010) Combinatorial alanine substitution enables rapid
optimization of cytochrome P450BM3 for selective hydroxylation of large
substrates, Chembiochem 11, 2502-2505.
Chapter 1
151.
152.
153.
154.
155.
156.
157.
158.
159.
160.
161.
162.
163.
164.
General Introduction
Weber, E., Seifert, A., Antonovici, M., Geinitz, C., Pleiss, J., and Urlacher, V. B.
Screening of a minimal enriched P450 BM3 mutant library for hydroxylation of
cyclic and acyclic alkanes, Chem Commun (Camb) 47, 944-946.
Sulistyaningdyah, W. T., Ogawa, J., Li, Q. S., Maeda, C., Yano, Y., Schmid, R. D.,
and Shimizu, S. (2005) Hydroxylation activity of P450 BM-3 mutant F87V towards
aromatic compounds and its application to the synthesis of hydroquinone
derivatives from phenolic compounds, Appl Microbiol Biotechnol 67, 556-562.
Misawa, N., Nodate, M., Otomatsu, T., Shimizu, K., Kaido, C., Kikuta, M., Ideno,
A., Ikenaga, H., Ogawa, J., Shimizu, S., and Shindo, K. (2011) Bioconversion of
substituted naphthalenes and beta-eudesmol with the cytochrome P450 BM3
variant F87V, Appl Microbiol Biotechnol 90, 147-157.
Nazor, J., and Schwaneberg, U. (2006) Laboratory evolution of P450 BM-3 for
mediated electron transfer, Chembiochem 7, 638-644.
Fasan, R., Chen, M. M., Crook, N. C., and Arnold, F. H. (2007) Engineered alkanehydroxylating cytochrome P450(BM3) exhibiting nativelike catalytic properties,
Angew Chem Int Ed Engl 46, 8414-8418.
Peters, M. W., Meinhold, P., Glieder, A., and Arnold, F. H. (2003) Regio- and
enantioselective alkane hydroxylation with engineered cytochromes P450 BM-3, J
Am Chem Soc 125, 13442-13450.
Murataliev, M. B., Trinh, L. N., Moser, L. V., Bates, R. B., Feyereisen, R., and
Walker, F. A. (2004) Chimeragenesis of the fatty acid binding site of cytochrome
P450BM3. Replacement of residues 73-84 with the homologous residues from
the insect cytochrome P450 CYP4C7, Biochemistry 43, 1771-1780.
Huang, W. C., Cullis, P. M., Raven, E. L., and Roberts, G. C. (2011) Control of the
stereo-selectivity of styrene epoxidation by cytochrome P450 BM3 using
structure-based mutagenesis, Metallomics 3, 410-416.
Rea, V., Kolkman, A. J., Vottero, E., Stronks, E. J., Ampt, K. A., Honing, M.,
Vermeulen, N. P., Wijmenga, S. S., and Commandeur, J. N. (2012) Active site
substitution A82W improves the regioselectivity of steroid hydroxylation by
cytochrome P450 BM3 mutants as rationalized by spin relaxation nuclear
magnetic resonance studies, Biochemistry 51, 750-760.
Venkataraman, H., Beer, S. B., Bergen, L. A., Essen, N., Geerke, D. P.,
Vermeulen, N. P., and Commandeur, J. N. (2012) A single active site mutation
inverts stereoselectivity of 16-hydroxylation of testosterone catalyzed by
engineered cytochrome P450 BM3, Chembiochem 13, 520-523.
Novak, R. F., and Vatsis, K. P. (1982) 1H Fourier transform nuclear magnetic
resonance relaxation rate studies on the interaction of acetanilide with purified
isozymes of rabbit liver microsomal cytochrome P-450 and with cytochrome b5,
Mol Pharmacol 21, 701-709.
Solomon, I., and Bloembergen, N. (1956) Nuclear magnetic interactions in the HF
molecule, J Chem Phys 25, 261-266.
Regal, K. A., and Nelson, S. D. (2000) Orientation of caffeine within the active site
of human cytochrome P450 1A2 based on NMR longitudinal (T1) relaxation
measurements, Archives of biochemistry and biophysics 384, 47-58.
van de Straat, R., de Vries, J., de Boer, H. J., Vromans, R. M., and Vermeulen, N.
P. (1987) Relationship between paracetamol binding to and its oxidation by two
Chapter 1
165.
166.
167.
168.
169.
General Introduction
cytochromes P-450 isozymes--a proton nuclear magnetic resonance and
spectrophotometric study, Xenobiotica; the fate of foreign compounds in
biological systems 17, 1-9.
Poli-Scaife, S., Attias, R., Dansette, P. M., and Mansuy, D. (1997) The substrate
binding site of human liver cytochrome P450 2C9: an NMR study, Biochemistry
36, 12672-12682.
Modi, S., Paine, M. J., Sutcliffe, M. J., Lian, L. Y., Primrose, W. U., Wolf, C. R.,
and Roberts, G. C. (1996) A model for human cytochrome P450 2D6 based on
homology modeling and NMR studies of substrate binding, Biochemistry 35,
4540-4550.
Modi, S., Gilham, D. E., Sutcliffe, M. J., Lian, L. Y., Primrose, W. U., Wolf, C. R.,
and Roberts, G. C. (1997) 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine as a
substrate of cytochrome P450 2D6: allosteric effects of NADPH-cytochrome P450
reductase, Biochemistry 36, 4461-4470.
Smith, G., Modi, S., Pillai, I., Lian, L. Y., Sutcliffe, M. J., Pritchard, M. P.,
Friedberg, T., Roberts, G. C., and Wolf, C. R. (1998) Determinants of the
substrate specificity of human cytochrome P-450 CYP2D6: design and
construction of a mutant with testosterone hydroxylase activity, Biochem J 331 (
Pt 3), 783-792.
Keizers, P. H., de Graaf, C., de Kanter, F. J., Oostenbrink, C., Feenstra, K. A.,
Commandeur, J. N., and Vermeulen, N. P. (2005) Metabolic regio- and
stereoselectivity of cytochrome P450 2D6 towards 3,4-methylenedioxy-Nalkylamphetamines: in silico predictions and experimental validation, J Med Chem
48, 6117-6127