94
Insulin as a regulator of flavin-containing monooxygenase enzyme in streptozotocin-induced diabetic rats Ph.D. Thesis Tímea Borbás Semmelweis University Doctoral School of Pharmaceutical and Pharmacological Sciences Supervisor: Dr. Károly Tihanyi, C.Sc. Opponents: Dr. Miklós Tóth, D.Sc. Dr. Zsuzsanna Veres, D.Sc. Final Examination Committee: President: Dr. Krisztina Takács-Novák, D.Sc. Members: Dr. Imre Klebovich, D.Sc. Dr. Katalin Monostory, Ph.D. Budapest 2006.

Insulin as a regulator of flavin-containing monooxygenase

  • Upload
    others

  • View
    6

  • Download
    0

Embed Size (px)

Citation preview

Insulin as a regulator of flavin-containing monooxygenase enzyme in streptozotocin-induced

diabetic rats

Ph.D. Thesis

Tímea Borbás

Semmelweis University Doctoral School of Pharmaceutical and Pharmacological Sciences

Supervisor: Dr. Károly Tihanyi, C.Sc. Opponents: Dr. Miklós Tóth, D.Sc.

Dr. Zsuzsanna Veres, D.Sc.

Final Examination Committee: President: Dr. Krisztina Takács-Novák, D.Sc. Members: Dr. Imre Klebovich, D.Sc.

Dr. Katalin Monostory, Ph.D.

Budapest 2006.

TABLE OF CONTENTS

ABBREVIATIONS...................................................................................................................4

1 INTRODUCTION .....................................................................................................6 1.1 DISCOVERY OF FMO ENZYMES ............................................................................6 1.2 BIOCHEMICAL, CATALYTICAL AND STRUCTURAL PROPERTIES................7

1.2.1 Classification of the FMO enzyme ............................................................... 7 1.2.2 Unique biochemical properties .................................................................... 8 1.2.3 Approaches to distinguish between FMO- and CYP-mediated metabolism. 9 1.2.4 The mechanism of catalysis ........................................................................ 11 1.2.5 Structure: gene and protein........................................................................ 12

1.3 ISOFORMS: TISSUE-, SPECIES-, GENDER-, AGE- AND SUBSTRATE-SPECIFICITY.............................................................................................................16

1.4 CATALYSED REACTIONS .....................................................................................22 1.4.1 Substrates: endogenous and exogenous ..................................................... 22 1.4.2 Stereoselectivity .......................................................................................... 24 1.4.3 In vitro and in vivo probes of FMO............................................................ 24 1.4.4 Inhibitors .................................................................................................... 26

1.5 ROLES OF FMO ........................................................................................................27 1.5.1 Metabolism of endogenous compounds...................................................... 27 1.5.2 Xenobiotic metabolism: detoxication or bioactivation............................... 29 1.5.3 FMO in drug development.......................................................................... 30

1.6 REGULATION...........................................................................................................31 1.6.1 Genetic variation ........................................................................................ 31 1.6.2 Dietary, developmental, hormonal control ................................................ 33 1.6.3 Pathophysiological status: diabetes ........................................................... 35

2 RESEARCH OBJECTIVES...................................................................................39

3 MATERIALS AND METHODS............................................................................40 3.1 BIOCHEMICALS ......................................................................................................40 3.2 ANIMALS AND INDUCTION OF DIABETES .......................................................41 3.3 PREPARATION OF LIVER MICROSOMES...........................................................41 3.4 DETERMINATION OF CYTOCHROME CONTENT.............................................42 3.5 ENZYMATIC ASSAYS.............................................................................................42

3.5.1 FMO index reaction: Benzydamine N-oxygenation ................................... 43 3.5.2 CYP1A index reaction: Ethoxyresorufin O-deethylation ........................... 43 3.5.3 CYP2B/3A index reaction: Aminopyrine N-demethylation ........................ 44 3.5.4 CYP2E1 index reaction: p-Nitrophenol-hydroxylation.............................. 44 3.5.5 CYP3A index reaction: Testosterone-6β-hydroxylation............................. 45

3.6 DETERMINATION OF THE mRNA OF FMO1 AND FMO3 ISOFORMS ............45 3.6.1 Isolation of total RNA................................................................................. 45 3.6.2 Measurement of concentration and integrity of RNA................................. 46 3.6.3 Synthesis of cDNA ...................................................................................... 46 3.6.4 Gene expression analysis ........................................................................... 46 3.6.5 Data analysis .............................................................................................. 47

3.7 FMO-MEDIATED BIOTRANSFORMATION OF SOME CNS DRUGS................47 3.7.1 Tolperisone ................................................................................................. 47

2

3.7.2 Deprenyl, amphetamine, methamphetamine enantiomers.......................... 48

4 RESULTS.................................................................................................................50 4.1 DRUG METABOLIZING MICROSOMAL ENZYMES IN DIABETIC RATS ......50

4.1.1 Changes of physical and biochemical parameters in diabetic rats............ 50 4.1.2 The function of FMOs................................................................................. 52 4.1.3 The function of cytochromes: CYP and cytochrome b5 .............................. 58 4.1.4 Correlation analysis ................................................................................... 60

4.2 FMO-MEDIATED METABOLISM OF CNS DRUGS.............................................62 4.2.1 Tolperisone ................................................................................................. 62 4.2.2 Deprenyl, methamphetamine, amphetamine .............................................. 63

5 DISCUSSION...........................................................................................................64 5.1 CHANGES IN THE CYTOCHROME ENZYME SYSTEM: CYTOCHROME

P450 AND CYTOCHROME B5.................................................................................64 5.1.1 Hepatic cytochrome contents and CYP isoform activities.......................... 64

5.2 CHANGES IN THE FMO ENZYME SYSTEM........................................................65 5.2.1 Insulin as a regulator of FMO.................................................................... 65 5.2.2 Glucose as a marker for elevated FMO activity......................................... 67 5.2.3 Ketone bodies may have a role in FMO regulation ................................... 67 5.2.4 FMO correlates with cytochrome b5 enzyme in experimental diabetes ..... 68

5.3 THE PROPOSED INFLUENCE OF DIABETES INDUCED CHANGES OF FMO ACTIVITY ON TOLPERISONE, DEPRENYL, AMPHETAMINE, METHAMPHETAMINE METABOLISM ................................................................68

6 CONCLUSIONS......................................................................................................70

7 SUMMARY..............................................................................................................71

8 REFERENCES ........................................................................................................73

9 PUBLICATIONS.....................................................................................................90

10 ACKNOWLEDGEMENTS ....................................................................................94

3

ABBREVIATIONS AA amino acid ANOVA one-way analysis of variance Asn 120 asparagine 120 BG blood glucose concentration BOR benzamidoxime reductase BZY benzydamine BZY N-oxide benzydamine N-oxide cDNA complementary DNA CYP cytochrome P450 cyt.b5 cytochrome b5D50 50 mg/kg streptozotocin induced diabetic rats D70 70 mg/kg streptozotocin induced diabetic rats DAG diacyl-glycerol DIM 3,3’diindolylmethane DTT dithiotreitol EDTA ethylenediamine-tetraacetic acid FAD flavin adenine dinucleotide FADH2 reduced flavin adenine dinucleotide FADOH hydroxy flavin dinucleotide FADOOH hydroperoxy flavin dinucleotide FMO flavin-containing monooxygenase GAPDH glyceraldehyde-3-phosphate dehydrogenase GDT guanylyl diphosphate GSH reduced glutathione GSSG oxidized glutathione GLUT-2 glucose transporter 2 Grb2 an adaptor protein containing SH2 GTP guanylyl triphosphate HLM human liver microsomes HNF1α hepatic nuclear factor, homeodomain-containing factor HNF4α hepatic nuclear factor, orphan nuclear receptor I3C indole-3-carbinol ID70 insulin treated 70 mg/kg streptozotocin induced diabetic rats IDDM insulin-dependent diabetes mellitus IGF-1 insulin-like growth factor IND insulin treated non-diabetic rats IP3 inozitol 1,4,5-triphosphate IRS insulin receptor substrate LT 2,3-bis[3-indolylmethyl]indole, (linear tetramer) MAO-B monoamine oxidase type B MAP kinase mitogen activated protein kinase MMI methimazole MPDP+ 2,3-dihydropyridinium MPP+ 1-methyl-4-phenylpyridinium MPTP 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine MPTP N-oxide 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine N-oxide

4

MpTS methyl para-tolyl sulfide MpTSO methyl para-tolyl sulfoxide mRNA messenger ribonucleic acid NADPH reduced nicotinamide adenine dinucleotide phosphate NADP+ nicotinamide adenine dinucleotide phosphate NIDDM noninsulin dependent diabetes mellitus PDB ID protein data bank identification PCR polymerase chain reaction PIP2 phosphatidyl-inositol 4,5 biphosphate PKC proteinkinase C PLC-γ phospholipase C-γ PTP 4-phenyl-1,2,3,6-tetrahydropyridine ras guanine nucleotide-binding protein RIN RNA Integrity Number RNA ribonucleic acid RLM rat liver microsomes SH2 domain src homology domain 2 SNP single nucletide polymorphism SOS Ras-specific nucleotide exchange factor Src kinase a non-receptor tyrosine kinase STZ streptozotocin TCA trichloro acetic acid TMA trimethylamine TMA N-oxide trimethylamine N-oxide TMAuria trimethylaminuria Tris tris hydroxymethyl aminomethane YY1 Yin Yang-1 1get PDB ID of glutathione-reductase 1npx PDB ID of NADPH-peroxidase 1vqw PDB ID of a protein with similarity to FMO 1w4x PDB ID of phenylacetone monooxygenase 5-DPT 10-([N,N-dimethylamino]alkyl)-2-(trifluoromethyl)phenothiazine

5

1 INTRODUCTION Drug metabolizing enzymes being adaptive enzymes are of great importance in

maintaining homeostasis. The flavin-containing monooxygenase (FMO) family is one

of the major microsomal monooxygenase enzyme systems involved in drug metabolism.

Its function is NADPH- and O2-dependent. FMO catalyses the oxygenation of a variety

of nucleophilic heteroatom-containing (i.e. nitrogen, sulfur, selenium and phosphorous)

xenobiotics to their respective oxides.1

FMO was purified from pig liver in Ziegler’s laboratory in 1972, hence it was

named „Ziegler’s-enzyme” for a period of time.2 Its contribution to drug metabolism

was underestimated until the 1990s. It is not inducible by typical cytochrome P450

(CYP) inducers, but in certain patophysiological states such as diabetes it is altered

along with CYP isoforms. In diabetes therapy, besides the description of metabolic

pathways, the recognition of changes in drug metabolising capacity is essential for

avoidance of unwanted drug-drug interactions and achievement of optimal drug

exposition. The regulatory role of insulin was studied regarding CYP isoforms, but

similar observations have not yet been made for FMO.

In this thesis, first the recent developments on flavin-containing monooxygenase

will be surveyed concerning its biochemical, structural and catalytical properties;

isoforms; substrate specificity; physiological significance and toxicological importance;

relevance in drug metabolism and development; regulation; functional SNPs causing

FMO deficiency leading to a hereditary disease, trimethylaminuria. Then, it will be

described how we came to the conclusion that insulin is a regulator of FMO.

1.1 DISCOVERY OF FMO ENZYMES

In 1960 Miller and his co-workers reported that liver microsomes contained enzymes

that catalysed the NADPH- and O2-dependent oxidation of butter yellow and a number

of other azo dyes.3 Ziegler studied the nature of enzymes catalysing these so-called

mixed function oxidation reactions. They selected the N-oxidation of N,N-

dimethylaniline as a model reaction since its N-oxide was easily detectable by a rapid

colorimetric measurement. They characterized the enzyme in liver microsomes. The

highest activity among a wide variety of vertebrates was consistently observed in

6

microsomes isolated from pig liver. One of Ziegler’s students, Caroline Mitchell,

succeeded in purifying the enzyme that catalyses the NADPH- and O2-dependent N-

oxidation of N,N-dimethylaniline from this source in 1972.2 The purified flavoprotein

was free of metals and other nonprotein components other than FAD or lipid.

This enzyme had been known as a mixed-function amine N-oxidase until 1979.

However, shortly after the isolation of the enzyme from pig liver Jollow and Cook

demonstrated that it catalyzed oxygenation of an exceptionally wide range of

xenobiotics that had no common structural features.4 The list includes inorganic and

organic compounds and some of the better substrates contain functional groups bearing

sulfur or selenium instead of nitrogen. Since the name “mixed-function amine N-

oxidase” or simply “N-oxidase” was restrictive, the flavoprotein was given the trivial

name flavin-containing monooxygenase, usually abbreviated as FMO.

1.2 BIOCHEMICAL, CATALYTICAL AND STRUCTURAL PROPERTIES

1.2.1 Classification of the FMO enzyme

FMO is a microsomal monooxygenase enzyme catalysing xenobiotic oxidation. Flavin-

monooxygenase and cytochrome P450 share a number of similarities with respect to

tissue, cellular and organelle expression, the utilization of oxygen and NADPH as

cofactors, and substrate- and metabolite-specificity. Enzymes involved in drug

oxidation5 are summarized in Fig. 1.

7

Xenobiotic oxidation

Non-microsomal oxidation Microsomal oxidation

Heme-containing Non-heme containig

Cytochrome P450(CYP)

Flavin-containingmonooxygenase

(FMO)

Alcohol-dehydrogenaseAldehyde-dehydrogenaseDihydrodiol-dehydrogenaseMonoamine-oxidaseSemicarbazide-sensitive amine oxidaseEnzymes of peroxisomal β-oxidationMolybdenum-containing monooxygenases(aldehyde-oxidase and xantine-oxidase)

Xenobiotic oxidation

Non-microsomal oxidation Microsomal oxidation

Heme-containing Non-heme containig

Cytochrome P450(CYP)

Alcohol-dehydrogenaseAldehyde-dehydrogenaseDihydrodiol-dehydrogenaseMonoamine-oxidaseSemicarbazide-sensitive amine oxidaseEnzymes of peroxisomal β-oxidationMolybdenum-containing monooxygenases(aldehyde-oxidase and xantine-oxidase)

Flavin-containingmonooxygenase

(FMO)

Fig. 1. Enzymes involved in the oxidation of xenobiotics

1.2.2 Unique biochemical properties

The following properties are unique to the FMO class of monooxygenases:

a) Relative thermal lability:

FMOs (except FMO2) usually are extraordinarily sensitive to heat. In the absence of

NADPH, about 85 % of the activity of most FMOs is lost if the tissue is left

standing at 45-55 °C for 1-4 minutes. NADPH prevents activity loss caused by heat,

therefore stabilizes FMO.6

b) pH dependency:

The optimal pH for FMO enzyme function varies among species (rabbit, mouse

FMO2 pH 10; pig FMO1 pH 8.3-8.5; mouse liver FMO3 pH 8.8-9.2), each

definitely higher than that of CYP isoforms (pH 7.4). However, it is not known if

the enzyme works more efficiently at higher pH because the pKa of most FMO

substrates is also about pH 8.5-10.7

c) Detergent dependency:

The activity of FMO depends on detergent concentration. The FMO activity

increases at concentrations of Triton X-100 between 0.05-0.5 % while above 0.5 %

it decreases.8 FMO1 and FMO3 are inhibited at low concentrations of anionic

8

detergents such as sodium cholate or fatty acids while FMO2 function at the sodium

cholate concentration as high as 1 % is not inhibited.9

d) Formation of relatively stabile hydroperoxy flavin intermediate:

Hydroperoxy flavin is protected from decomposition by the protein environment of

FMO, and the presence of NADPH which is an FMO stabilizer.10 This allows FMO

to be in an oxygenated, so-called “cocked gun” mode. In the absence of substrate,

FMO does not generate copious amounts of H2O2 suggesting that FMO does not

expose the cell to untoward effects of oxidative stress.6

e) Metabolite prediction:

The metabolite produced by FMO in case of a particular substrate could be

predicted in case the substrate reacts with alkyl hydroperoxides or peracids.10

1.2.3 Approaches to distinguish between FMO- and CYP-mediated metabolism

FMO and CYP have overlapping substrate-specificity. During drug development it is

important to identify metabolic pathways, namely specific human oxidative enzymes

responsible for biotransformation. Such information can facilitate the rational,

prospective evaluation of drug-drug interactions and improve the safety profile of many

substances, particularly those with low therapeutic indices.11

In order to determine the relative contribution of FMO versus CYP in the

oxygenation of a particular drug, a number of in vitro approaches have been used.

a) Thermal inactivation:

Heating the enzyme up to 50 °C for 1 minute the activity is lost. FMO is relatively

unstable at this temperature, while 85 % of CYP functional activity remains.12

b) pH of incubation mixture:

Running microsomal incubations at pH 8.4-9.4 generally abrogates CYP activity.12

c) Detergents:

In the presence of high concentration of Emulgent911 CYP activity is abrogated.12

9

d) Metal ions:

FMO is relatively sensitive to metal ions, therefore incubations should be carried out

in the absence of MgCl2 and in the presence of a chelating agent.10

e) Inactivation of heme prostetic group via saturating by CO.9

f) Inhibitors:

Chemical inhibitors:

o FMO inhibitors: methimazol (MMI)13, thiourea14, indole-3-carbinol

(I3C).15

o General CYP inhibitors: N-benzylimidazole13, aminobenzotriazole,

proadifen hydrochloride (SKF-525).16

Immunological inhibitors (antibodies):

o FMO inhibitors: antibodies directed toward FMOs (rabbit lung

FMO217 and human liver FMO318).

o General CYP inhibitor: antibody against NADPH-P450 reductase.19

g) Molecular biology tools: recombinant FMO isozymes.9

h) Banks of human liver microsomes: well-characterized for CYP and FMO

activities in order to carry out correlation studies.9

i) Observing the stereochemistry of the product: the stereochemistry of FMO- and

CYP-mediated N- and S-oxygenations is sometimes distinct.10

It is more difficult to determine a metabolic pathway in vivo. In animals, pretreatment

with MMI or I3C inhibits FMO mediated metabolism, but these are not fully specific

for FMO. MMI inhibits thyroid peroxidase20 and the reactive sulfenic acid produced

inhibits CYP activity as well21, whereas I3C induces CYP isozymes22.

10

1.2.4 The mechanism of catalysis

Both, FMO and CYP require NADPH and O2 as cofactors for their catalytic activity.

The mechanism of reactions catalysed by FMO is different from that of the CYP-

mediated metabolism. In contrast to CYP, FMO binds the substrate only when its FAD

component has undergone NADPH-mediated two electron reduction, and the reduced

flavin reacted with molecular oxygen to form 4α-hydroxyperoxyflavin to be ready for

oxygenating any soft nucleophilic xenobiotics of which shape, size and charge permits

its access to the well-protected substrate binding channel. The intermediate is relatively

stabile and in this state like a “cocked gun” waits for the suitable substrate. Oxygenation

proceeds through an attack of the nitrogen atom (or any other nucleophilic heteroatom)

on the terminal hydroperoxy flavin oxygen atom to produce the N- or S-oxygenated

substrate and the hydroxy flavoenzyme species. Thus, one atom of molecular oxygen is

transferred to the substrate and the second to form water. The rate limiting step of the

catalytic cycle is thought to be the breakdown of the FADOH pseudobase or the release

of NADP+. As it occurs after the substrate transformation, substrate binding has no

influence on Vmax.6,9 The catalytic cycle is summarized in Fig. 2.

N

NN

N

R

O

O

H

N

NN

N

R

O

O

H

H

H

NADP

N

NN

N

R

O

O

H

H+H+

HO

O

H

O2

O

HH

N

NN

N

R

O

O

H

NADP++H2O

S

S

FAD

FADH2

FADOOH

FADOH

O

Fig. 2. The mechanism of FMO catalysis

The red arrow represents the rate-limiting step of catalysis. S indicates the nucleophilic heteroatom of the substrate.6

11

1.2.5 Structure: gene and protein

Currently eleven human FMO genes are known. FMO1, 2, 3, 4 and 5 isoforms are

functionally active the others are pseudogenes.23,24 The FMO gene family probably

arose by duplication of a common ancestral gene some 250-300 million years ago.25

The genes are located on the long arm of chromosome 1.26 The structural

organization of the human FMO3 gene is seen in Fig. 3.

Fig. 3. The gene structure of human FMO3 The exon/intron structure of the gene along with the arrangement of the exon-derived sequences within the corresponding mRNA are shown above. Introns (horizontal open boxes) are numbered in boldface type and their approximate sizes (in kilobases) are shown in paretheses. Exons (solid vertical boxes) are numbered in italics and are linked by dashed lines to the equivalent regions within the mRNA (open boxes) that contain the exon length (in nucleotides) or, in the case of exons 2 and 9, the length of the protein coding regions within the exon. Shaded regions represent 5’ and 3’ untranslated regions within the mRNA. The location within the mRNA of the sequences that encode the fingerprint motifs associated with the ADP-binding βαβ-folds of the FAD- and NADP-binding domains within the protein are shown by filled and open horizontal boxes, respectively.27

The deduced amino acid sequence of the human FMO isoenzymes have 82-87 %

identity with their known orthologues in other mammals, but only 51-57 % similarity to

each other25 with the exception of FMO3 and FMO6, which share 71 % identity28. FMO

human genes having sequence identity of ≥ 82 % are grouped in one family, which is

indicated by the first numeral of the designation (i.e. 1, 2, 3, 4, 5 and 6). The order of

12

naming followed the chronology of publication of the sequence for each member of the

family.27

The molecular weight of FMO is around 56 kDa.29 FMOs are built of 532-558

amino acids (AA) with highly conserved regions corresponding to the binding site of

the ADP moiety of FAD (AA position 4-32) and NADPH (AA position 186-213). FMO

apoenzyme binds FAD stochiometrically. All mammalian FMOs are anchored proteins

and possess very strong membrane association properties.6 Only its C-terminus

(residues 510-533) is sufficiently hydrophobic to be inserted into the membrane.

Instruction for active membrane association is probably encoded in an internal sequence

(residues 55-77). It was proposed that the substrate binding site of FMO3 is between the

397 and 431 amino acids. Ziegler has suggested that only a single point of attachment to

the terminal hydroperoxy flavin oxygen is required for substrate oxygenation. In

Cashman’s opinion, however, additional points of contact with the substrates are

required at the active site of FMO for explaining stereoselectivity. Other regions of

FMO have notable homology with well-characterized enzymes such as the serine

protease, acetyl-choline esterase, esterase and the leucin-box of T cell receptor.7

Since FMO is a membrane associated protein, its x-ray structure has not yet been

solved, however, the molecular models of FMO, based on the crystal structure of other

flavoproteins, have been proposed. The first FMO model was developed by Ziegler

based on the crystal structure of E. coli glutathione reductase.30 Ziegler suggested that

the FMO protein ought to be a dimer, since it seems to belong to the flavocytochrome c

sulfide dehydrogenase subfamily of flavoproteins that were described to be active in

their dimeric form (Fig. 4.).31 In a different approach, developed by Cashman and

Adman, the structure of NADPH-peroxidase was used to model human FMO3.6 A third

model of human FMO3 was built using four PDB structures (glutathione-reductase,

1get; NADPH-peroxidase, 1npx; a protein with similarity to flavin-containing

monooxygenase, 1vqw and phenylacetone monooxygenase, 1w4x) based on homology

modeling. The structure of human FMO3 and the distribution of structural domains9on

primary structure of human FMO3 is depicted in Fig. 5. (Borbás T, Zhang J, Cerny AM,

Likó I, Cashman JR, [Epub ahead of print]).

13

Fig. 4. Postulated tertiary ribbon structure of the human FMO3 dimer

The model was generated by replacing all the amino acids in E. coli glutathione reductase with those in the same position in FMO3.31

14

Fig. 5. The human FMO3 tertiary structure and distribution of structural domains

on primary structure of human FMO3 The model structure of human FMO3 above shows the highly conserved FAD (space filling representation colored by magenta) binding domain (red ribbon) and the NADPH (space filling representation colored by purple) binding domain (yellow ribbon). The FMO specific putative substrate binding region is shown with blue ribbon. The figure below shows the highly conserved FAD binding domain, the NADPH binding domain, the FMO specific putative substrate binding region and the putative membrane anchor region coloured with red, yellow, blue and green, respectively.

15

1.3 ISOFORMS: TISSUE-, SPECIES-, GENDER-, AGE- AND SUBSTRATE-SPECIFICITY

FMO isoforms exhibit tissue-, species-, gender-, age- and substrate-specificity.

a) Tissue-specificity: (Fig. 6.)

FMOs are expressed in the liver – the main metabolic organ – and in the lung, kidney,

small intestine, brain as well. FMO isoforms show the following tissue-specific

distribution:

FMO1 isoform is the most abundant in human kidney; human small intestine,

fetal liver, most experimental animal liver, nasal mucosa and esophagus also

contain it.

FMO2 isoform is dominant in human and rabbit lung, however it is not a

prominent active enzyme there.

FMO3 isoform is the most abundant in adult human liver, with a concentration

of 100 pmol/mg microsomal protein.32 The microsomal FMO3 content is 60 %

of CYP3A4 - the most abundant hepatic CYP isoenzyme – present in adult

human liver. Its turnover number is 2-3-fold greater than that of CYP.33,34

FMO4 isoform is dominant in adult human liver and kidney.

FMO5 isoform was shown to be as much abundant in adult human liver as

FMO3.

In humans, FMO is expressed in the highest quantity in the liver and lung and only

half of those amounts in the kidney. The abundance of various FMO isoforms in the

brain are represented equally less than 1 % in comparison to the richest corresponding

human tissues.

It is important to keep in mind when interpolating liver microsomal data for FMO

catalysis from experimental animals to humans that in animals the prominent hepatic

enzyme is FMO1, whereas in humans FMO3.5

16

Fig. 6. Tissue-specific distribution of FMO isoforms in humans Data based on mRNA measurements.35

b) Species-specificity: (Table 1.)

The pattern of tissue-specific distribution depends on the species examined. The table

shows a compilation of the „best guess” as to the tissue distribution of FMO forms in

experimental male animals and humans.

17

Table 1. Tissue levels of FMO forms in animals and humans

Tissues species FMO1 FMO2 FMO3 FMO4 FMO5 mouse low NP high ? high

rat high ? low ? low rabbit high NP low ? low

Liver

human very low low high high high mouse high ? high ? low

rat high ? high high low rabbit low low very low high low

Kidney

human high low ? high ? mouse ? high very low NP low

rat ? ? ? NP low rabbit ? very high ? NP NP

Lung

human ? high ? NP ?

NP, apparently not present. The question mark indicates that no data are available or the presence of an FMO form is in doubt. It is largely based on mRNA data.

Using a combination of microsomal activity measurements and immunoblot techniques

it was shown that the order of FMO3 levels is as follows:

rabbit>>mice=dog=human>rat. From the aspect of human hepatic metabolism, rat is a

poor model insofar FMO3 is concerned, because the prominent FMO present in rat liver

is FMO1. Female mice and dogs would be prudent to utilise.36

c) Gender-specificity: (Table 2.)

It was also shown that the expression of FMO3 is affected by gender. In mice and rats a

remarkable gender effect was observed with females having much greater activity than

males. While testosterone repressed, estradiol increased the expression of mouse FMO1

and FMO3. No gender effect was observed in humans, dogs and rabbits.37,38 The

human, rat and mouse gender-dependent hepatic FMO expression determined by

immunoreactivity assay is shown in Table 2.38

18

Table 2. Gender-specificity of FMOs

species gender FMO1 FMO3 male + + rat female - + male - - mouse female + + male N.D. + human female - +

Hepatic FMO1 is gender-dependent in rat and mouse, selective to male rat and female

mouse. Human FMO1 is nearly undetectable. Hepatic FMO3 in mouse is gender-specific to the female, but gender-independent in rat and man. N.D. indicates non

detectable quantity, + indicates gender-specific expression and – indicates it is not expressed.38

d) Age-specificity:

It was shown that the expression of human brain FMO1 was down-regulated during or

shortly after birth.35 Due to a developmental control FMO1 and FMO3 change in the

opposite direction in both human and mouse liver. This statement is going to be

discussed in more details in chapter 1.6.2.

e) Substrate-specificity:

Numerous structure-activity studies with purified and microsomal flavin-containing

monooxygenases suggest that the overall size, shape and charge of nucleophilic

xenobiotics are the major factors that limit access to the hydroperoxyflavin of FMO and

these are responsible for the differences in the specificities of the isoforms.31,39

Structural requirements for FMO substrates are:

1. Any compound containing a soft nucleophil that is accessible to the FADOOH

can become a substrate.

2. Compunds containing a single positive charge are excellent substrates (e.g.

cationic tertiary amines).

3. Negatively charged compounds are excluded entirely or are poor substrates with

a few exceptions, such as sulindac sulfide and lipoic acid.

4. Zwitterions and compounds with more than one positive charge are typically not

substrates.9

19

In general, FMOs have broad and overlapping substrate-specificity. FMO1, 2 and 3

oxygenate a wide variety of nucleophilic tertiary and secondary amines as well as

sulfur-containing compounds compared to FMO4 and 5. FMO5 has restricted substrate-

specificity, presumably since it does not form a stable intermediate.31 The substrate-

specificity of FMO5 is poorly defined although is apparently distinct from FMO3.

Because of their limited substrate specificity and low expression levels in most tissues,

FMO4 and FMO5 currently are not thought to play an important role in drug

metabolism.9,40

The main properties regarding substrate-specificity of FMO isoforms are

summarized in Table 3.

20

T

able

3. T

he su

bstr

ate-

spec

ifici

ty o

f FM

O is

ofor

ms

The

subs

trat

e-sp

ecifi

city

of F

MO

isof

orm

s ,,

,,

C

PZ: c

hlor

prom

azin

e, IP

M: i

mip

ram

ine

21

1.4 CATALYSED REACTIONS

1.4.1 Substrates: endogenous and exogenous

FMO catalyses the oxygenation of nucleophilic heteroatom-containing (i.e. nitrogen,

sulfur, selenium and phosphorous) substrates. Sulfur atom is a preferred site of FMO

oxygenation owing to its enhanced nucleophilicity, therefore S-oxygenation is favored

over N-oxygenation.10

a) Endogenous substrates

As discussed previously, the charge restriction for access to the substrate channel

leading to FADOOH exclude many potential nucleophilic endogenous substrates from

FMO-dependent oxygenation. There are a number of notable exceptions, however.

Substrates and their metabolites produced by FMO are listed below:

Nitrogen containing:

o biogen amines (phenethylamine, tyramine) → trans-oxime41

o trimethylamine → trimethylamine N-oxide, KM=28 uM10

Sulfur-containing:

o Cysteamine (sulfhydryl) → cystamine (disulfide), KM=120 uM

o disulfide lipoic acid → sulfoxide, KM=120 uM

o methionine → sulfoxide, KM=20 mM

o S-farnezyl-cysteine → sulfoxide, KM=30 uM

o cysteine S-conjugates, selenocysteine conjugates with high KM9

22

b) Exogenous substrates

There are several pharmaceutical agents for which FMO has been shown to be the

primary determinant of efficacy/toxicity.

Nitrogen-containing:

o Amines9,10,42

Primary alkyl amines N-hydroxylamine oxime

Secondary amines N-hydroxylamine nitrone hydroxylated primary amine+ aldehyde or ketone

Tertiary amines FMO

CYP orreductases

N-oxide

FMO

FMOBOR

BORFMO

FMO

hydrolysis-H2O

-H2O

o 1,1 disubstituated hydrazine → aldehyde + dealkylated hydrazine

o Heterocyclic amines: compounds containing nucleophilic cyclic tertiary amines

such as N-methyl tetrahydropyridines, piperidines, piperazines and pyrrolidines

are the best substrates for FMO3. The most likely reason for this is the enhanced

nucleophilicity of the N-atom.36

Drugs: chloro- and bromo-pheniramine, zimeldine, ranitidine, benzydamine,

olopatadine, xanomeline, pargyline, itopride.

Sulfur-containing:

o thiols → [sulfenic acid] → disulfides → sulfoxides

o sulfides → sulfoxides → sulfones

o thiones (thiobenzamide, thiocarbamate, thiourea) → sulfines, sulfenes →

sulfates

o cystein S-conjugates → sulfoxides

o The best substrate is tetrahydrothiophene, a cyclic sulfide with sub-micromolar

KM.

Drugs: cimetidine, albendazole, sulindac sulfide, methimazole, ethionamide.9

23

1.4.2 Stereoselectivity

FMO-mediated oxygenation often shows great stereoselectivity.

Primary alkyl amines are transformed to oximes in a stereospecific fashion.

Trans-oxime is formed from amphetamine, tyramine41, phenetylamine43, while cis-

oxime is formed from 5-DPT41.

Chiral sulfides are transformed to sulfoxide in a stereospecific fashion.

Cimetidine36, parylgine25, ,44 and flosequinan45 are transformed into (-) S-oxide with

FMO3 and (+) S-oxide with FMO1. Methionine is transformed into d-methionine S-

oxide.46

Often, the contribution of FMO to the metabolism of a given compound can be

assessed by its unique stereoselectivity relative to other oxygenases. For example, the

cytochrome P450s oxidize (S)-nicotine to a mixture of cis- and trans-N-1'-oxides. In

contrast, (S)-nicotine is oxidized by human FMO3 exclusively to the trans-N-1'-oxide.47

1.4.3 In vitro and in vivo probes of FMO

In order to determine FMO activity in vitro and in vivo FMO-specific substrates are

needed. There are two substrates commonly used to measure both, in vivo and in vitro

hepatic FMO activity.

Trimethylamine (TMA) is an endogenous substrate; it can be used in a non-

invasive-way to quantify human FMO phenotype. TMA/TMA N-oxide transformation

could be considered as a reaction catalysed exclusively by FMO3 since human FMO3 is

at least 100-fold more efficient at oxygenating TMA than human FMO1. The

extrahepatic human FMO1 probably does not make a significant contribution to TMA

N-oxygenation.48 The ratio of TMA and TMA N-oxide in urine is a biomarker of human

hepatic FMO3 activity.

Benzydamine (BZY) is transformed into BZY N-oxide and desmethyl-BZY with

liver microsomes. BZY N-oxide, the major metabolite, is produced mainly by FMO as

it was proved in a study with human CYP and FMO recombinant proteins.49 Desmethyl-

BZY, the minor metabolite, is formed by CYP (Fig. 7.).50

24

Fig. 7. Structure of benzydamine and pathways of benzydamine metabolism in liver microsomes

Benzydamine N-oxygenation is the major pathway compared to benzydamine N-demethylation. ,51 Low levels of N-oxide are also formed by CYP isoforms.

In humans, benzydamine is metabolised mainly to N-oxide (90 %) and only 10 % of the

dose administered is N-demethylated by CYP3A4.52 N-oxygenation is catalysed

primarily by FMO3, however its KM is lower and Vmax is higher for FMO1 (Table 4.).

As FMO3 is the dominant isoform in human liver, FMO1 can contribute only to the

extrahepatic metabolism. BZY is a good FMO specific in vitro probe and also an easy-

to handle human diagnostic probe for FMO3 deficiency.51

Table 4. KM, Vmax and catalytic efficacy values for FMO isoforms with BZY

Cell fraction KM(uM)

Vmax(nmol*mg-1*min-1)

Catalytic efficacy (ml*min-1*mg-1)

FMO1 23.6 40.8 1.7 recombinant FMO3 40.4 29.1 0.7

purified FMO1 15 1870 125 Human 64 6.9 0.1 liver

microsomes Rat 15 37 2.5

Parameters were determined in phosphate buffer at pH 7.4-7.6. ,51

25

S-nicotine would be a highly stereoselective probe of human FMO activity, but

because of its relatively high KM value, its usefulness as an in vivo probe is less than

desired. Since human FMO3 produce exclusively trans-nicotine N-oxide, whereas

animal FMO1 produces cis and trans nicotine N-oxide, S-nicotine is reasonably useful

as a high stereoselective in vitro probe of human FMO3.36

There are some probe drugs that are used mainly in vivo for population studies.

Cimetidine, mainly in (-) form, transformed to cimetidine S-oxide in a ratio of (-) 75 %

and (+) 25 % in urine. Clozapine was used to analyse population frequencies and allelic

linkage of human FMO3 variations.36 Ranitidine was used to phenotype Korean

population and correlate with human FMO3 genotype.53

1.4.4 Inhibitors

FMO has a few competitive and uncompetitive inhibitors, but to date no mechanism-

based (suicide substrate) inhibitor is known. The most commonly used competitive

inhibitors of FMO in vitro are MMI, thiobenzamide and I3C.6,13,22

The mixture of I3C gastric acid condensation products were acting as competitive

inhibitors, both in vitro and in vivo. In vivo I3C is an autolysis product of 3-indole-

methyl-glucosinolate that presented in Brussels sprouts. I3C is not stable after transit

through the stomach, and it forms gastric acid condensation products. In rats fed with

I3C, CYP1A1 is induced more than 20-fold, whereas modest increases of 2 to 4-fold

were observed for CYP1A2, CYP2B1/2 and CYP3A simultaneously to a markedly

reduced FMO-catalysed reaction.22 In guinea pig CYP was induced, but FMO was

unchanged. Hence, it was concluded that CYP1A1 induction and FMO repression by

dietary I3C occur through independent mechanisms. I3C acid condensation products are

capable of binding to the aryl-hydrocarbon receptor (AhR) and eliciting CYP1A

induction.54 In vitro and in vivo studies were conducted in order to evaluate the

inhibitory capacity of I3C derivates. Ki values observed were in the interval that

apparently physiologically relevant. It was shown that two condensation products,

namely 3,3’diindolylmethane (DIM) and 2,3-bis[3-indolylmethyl]indole (LT), are

largely responsible for the human FMO3 inhibitory activity.55 Clinical and experimental

studies indicated that the inhibition effect of I3C gastric condensation products is

26

species-specific. In humans, rats, mice, rabbits and guinea pigs the remaining activities

were 30, 30, 100, 60 and 100 %, respectively.15,22,55,56

Dimethylamino stilbene carboxylate is an uncompetitive inhibitor of FMO, which

appears to interfere with the NADPH-binding domain.10 Polyclonal antibodies against

FMO are reportedly less successfully used.9

1.5 ROLES OF FMO

FMO is present both in prokaryotes and eukaryotes.57 It is believed that FMO evolved

to protect mammals from an onslaught of lipophilic and nucleophilic chemicals in the

early environment.58 In plants, namely Arabidopsis, an FMO-like enzyme called

YUCCA enzyme catalyses a key step in tryptophan-dependent auxin biosynthesis.59

It is known that yeast FMO (yFMO) generates oxidizing equivalents in the form

of GSSG that are then transported into the lumen of the endoplasmic reticulum and

create a redox potential essential for proper folding of proteins that have disulfide

bonds.60 However, yFMO differs from the mammalian FMO as the former one easily

accepts substrates like glutathione, cysteine and cysteamine.61

The importance of mammalian FMO in xenobiotic metabolism is proved, while

its physiological role has not been identified yet, except for TMA N-oxygenation.9

1.5.1 Metabolism of endogenous compounds

FMO is responsible for transformation of the odorous trimethylamine (TMA) into

odourless TMA N-oxide. TMA is derived from the diet by enterobacterial

decomposition of precursors such as TMA N-oxide, choline, lecithine and carnitine.

TMA N-oxide is readily excreted by the urine.27 The failure to efficiently oxygenate

TMA leads to trimethylaminuria (Fish Odour Syndrome).

Endogenous compounds, such as biogen amines (phenetylamine, tyramine) are

good substrates for FMO. The oximes appear to have little pharmacological activity,

and therefore FMO-mediated N-oxygenation of biogenic amines is belived to be one of

the mechanisms of inactivation.9

Farnesyl-cystein and farnesyl-cystein methyl ester are efficiently S-oxygenated

in a stereoselective fashion by FMO (KM=22.7 and 33.9 uM, Vmax=1319 and

27

1001 nmol*min-1*mg protein-1, respectively). Metabolising the terminal cysteine of the

farnesylated ras protein may prevent that ras protein associates with the membrane and

eventually MAP kinase cascade activation does not happen. Further studies are required

to establish this possible role.7

It is known that mammalian FMO strongly prefers not to accept physiological

nucleophiles (i.e. gluthatione and cysteine) that would otherwise cause a nonproductive

futile cycle of NADPH consumption and glutathione depletion. As the KM values for

methionine and cysteine S-conjugates are at mM concentration, it may call into question

the importance of such metabolism in vivo.9,40

The physiological role of S-oxygenation of cysteamine to cystamine by FMO is

not known. It was suggested that FMO has a role in protein disulfide bond formation

through oxygenating of cysteamine.62 Based entirely on speculation it is thought that

FMO oxygenation of cysteamine may serve to help control the overall thiol/disulfide

redox state of the cell, which in turn, controls many important pathways. FMO could

have a simple protective function. Cysteamine above the concentration of 40 uM is

toxic to cells, apparently through the transition metal-dependent formation of H2O2.63

Therefore, FMO may control the H2 2

2 2 Cysteamine can reach mM

cellular concentration. Hence, FMO oxygenation of cysteamine to cystamine followed

by its transport to the cytosol may represent a detoxication mechanism.

O level and the expression of genes regulated by

H O , sulfhydryl/disulfid ratios and general redox state.64

FMO catalyzes the sulfoxydation of the methyl sulfids of lipoic acid, however its

potential role in lipoic acid metabolism is not yet known. Lipoic acid plays important

role in energy metabolism (i.e. cofactor of the piruvate dehydrogenase enzyme

complex) and protection of mitochondria during aging.65

Reactive metabolites formed by FMO have not yet been observed to inactivate

the FMO enzyme, however they inhibit or covalently modify proximal proteins,

including CYPs.21,66

28

1.5.2 Xenobiotic metabolism: detoxication or bioactivation

a) Detoxication

FMO is involved in the oxygenation of a wide range of heteroatom-containing

compounds transforming xenobiotics to more polar, readily excreted (N-oxide or S-

oxide) forms. Generally, FMO takes part in the detoxication of substrates such as

synthetic therapeutic drugs and natural compounds in the diet like alkaloids. The N-

oxygenation of tertiary amine usually results in detoxication.

E.g.:

MPTP

FMO

MAO-B

MPTP N-oxide: readily excreted

MPDP+ MPP+: resulting parkinsonism

CYPPTP

FMO-mediated metabolism might be underestimated since tertiary amine N-oxide

can be reduced back to its parent amine by CYPs or reductases (eg. nicotine,

tamoxifen), just as N-hydroxylamines to primary and secondary amines by

benzamidoxime reductase (BOR). By this mechanism FMO provides a reservoir of

parent drug, prolonging its action. Since FMO often contributes to the formation of the

same metabolite pool produced by CYPs, its action is frequently overlooked.9

Regarding sulfur-containing compounds FMO prevents other cellular

macromolecules by being damaged from reactive metabolites such as sulfines, sulfenes

and sulfenic acids generated by other enzymes like CYPs.10

b) Bioactivation

The FMO enzyme family has also been implicated in the bioactivation of several

xenobiotics resulting in metabolites with greater electrophilicity, as mentioned above

(i.e. sulfenic acid, sulfinic acid).7,40 Thiourea S-oxygenation is an example of FMO-

dependent bioactivation of xenobiotics. The sulfenic acid produced reversibly reacts

with GSH and drives oxidative stress through redox cycles; whereas sulfinic acid is

responsible for inactivating CYPs.7

The N-oxygenation of secondary amines does not always result in detoxication. An

FMO-mediated metabolism of N-deacetyl-ketoconazole appears to be involved in

29

hepatotoxicity. FMO-dependent production of hydroxylated secondary and primary

amines can lead to complexation with hemoproteins including CYP, resulting in

inhibition and toxicity.67 The adverse effect (hepatotoxicity) of ethionamide is due to

FMO mediated metabolism.9

1.5.3 FMO in drug development

Xenobiotic metabolism primarly occurs via CYPs. Hepatic CYP3A isoform has a

dominant role as it is responsible for transformation of 50-60 % of drugs metabolised by

CYPs, and as a consequence drug-drug interactions could mainly be associated to

CYP3A.

In drug development many candidates have failed on pharmacokinetic and

toxicological causes. For example, toxic metabolites or bioavailability problems were

responsible for many failures. The aim of drug research is to develop efficient and safe

drugs with acceptable bioavailability. Utilizing the drug metabolism pathway involving

flavin-containing monooxygenase and minimizing P450-mediated bioactivation could

mean a strategy which may produce metabolites with lower or no toxicity, lead to

higher bioavailability and decrease the chance of drug-drug interactions as the

metabolism of a chemical is distributed over a wider scale of monooxygenases.

a) Drug metabolism shifted toward FMO-mediated metabolism

Cisapride, mosapride and itopride are gastrokinetic agents. Cisapride and mosapride are

metabolized mainly by CYP3A4-mediated dealkylation. Itopride mainly undergoes

FMO-catalyzed tertiary amine N-oxygenation (75 % of oral dose in humans). The

adverse effect (life-threatening ventricular arrhythmias) observed with cisapride with

concurrent administration of CYP3A4 inhibitor was not seen using itopride.

b) Prodrug approach

Retroreduction of oxygenated N-containing compounds may provide a novel prodrug

approach. An enzyme system that requires NADH-cytochrome b5 reductase,

cytochrome b5 and a third component with the characteristic of cytochrome P450

(CYP2D6 or CYP2A) is responsible for hydroxylamine reduction. For drugs rapidly

30

metabolized or for those with problematic pharmacokinetic properties, utilizing a

strategy of preparing N-oxygenated prodrugs could improve drug-likeness. E.g.:

Benzamidine BenzamidoximeCYP/FMO

BOR Further examples are pentamidine, sibrafiban and melagatran.68

1.6 REGULATION

1.6.1 Genetic variation

The basal FMO activity is genetically determined. Primarily, genetic polymorphisms

are responsible for the interindividual variation of FMO-mediated endogenous substrate

and drug metabolism.

The FMO functional diversity is determined by expression of five FMO genes,

in which single nucleotide polymorphisms were shown. For FMO1 it does not have a

great importance because observed SNPs have similar activities to the wild-type. FMO4

and FMO5 have only a few SNPs discovered. There is a significant truncation mutation

for human FMO2. Individuals with mutant FMO2 may be more protected from the toxic

properties of thioether-containing pesticides, whereas humans with full-length FMO2

may be more susceptible to toxicity of certain thiourea-containing chemicals because of

the sulfenic acid produced. Known SNPs of FMO3 can cause an increased catalytic

efficiency (i.e. L360P) or a deficiency of FMO function. SNPs of FMO3 causing

minimal or no functional activity are: double mutant K158/G308 (prevalence 20%);

P153L, M66I and E305X (rare mutations). Additional trimethylaminuria-causing

mutations include mutations that result in frame shift and cause premature termination

of the FMO3 protein synthesis.

There are a large number of splice variants, yet full-length FMO is the

predominant form presented for all FMO isoforms except for FMO4.40

FMO3 deficiency caused by the genotypes mentioned above results in an

enzyme with lower capacity of transforming odorous TMA into odourless TMA N-

oxide. Therefore an excess of the extremely odourous TMA in the urine, sweat and

breath of such patients is detectable with a body odor resembling rotten fish. These

symptoms are accompanied by various phsyhosocial abnormalities.69, ,70 71 This

31

metabolic disease is called: trimethylaminuria. It was first mentioned in the

Mahabharata, in an Indian epic tale (1400 B.C.), where a young woman was ostracized

from the society because she smelled like rotting fish. Moreover, Shakespeare was

likely referring to a TMAuria patient in a scene from “The tempest”.34 The first clinical

description was given by Humbert in 1970, on a young patient.72

There are several subtypes of this disorder: 1. primary genetic (autosomal

recessive hereditary disease), 2. hepatitis induced form, 3. childhood transient form

(immaturity of the enzyme), 4. transient form associated with menstruation, 5. precursor

overload (Huntington chorea or Alzheimer treatment with choline at high daily dosage),

6. disease states (liver cirrhosis, uremic patients).34

The incidence of trimethylaminuria has been reported to be between 0.1-1% in

British Caucasians.73 A higher prevalence of trimethylaminuria is observed in the

tropics which maybe a protection from insects.74 The disease is diagnosed by urinary

ratios of TMA-NO and TMA following a challenge dose of choline. The effective

treatment appear to be dietary restriction (low choline, carnitine, lecithine, fish),

antibiotics to inhibit bacterial reduction of TMA N-oxide in the intestine (e.g.

metronidazole), laxatives to decrease transit time (e.g. lactulose), enchancement of

residual FMO3 activity by supplementation of riboflavin, folate supplementation since a

methyl donor may interact with choline and methionine in the metabolic processes,

charcoal and copper-chlorophyllin (deodorant). Individuals having functional

polymorphisms of FMO3 metabolise related substrates (dietary compounds, drugs and

pesticides) in a distinct manner.34

Several psychiatric disturbances have been reported in patients with TMAuria

and these have been presumed to be psychosocial reactions to the odour. Feelings of

shame, low selfesteem, social isolation, anxiety, depression, suicidal tendencies,

paranoia, educational and career disadvantages, relationship difficulties, addiction to

alcohol and other drugs and overt aggression have been described. Speculation of a

causative linkage between neurological disorders and deficient TMA N-oxygenation

could be conceivable along several lines.75

32

1.6.2 Dietary, developmental, hormonal control

In contrast to CYPs, FMO is not readily induced or inhibited by exogenous compounds.

Thus, changes in FMO activity as a consequence of environmental influences are not

likely to be dramatic.

FMO is regulated at transcriptional, posttranscriptional and posttranslational

levels. A number of factors can influence the expression of FMO (physiological and

dietary factors)40 and alter the activity of FMO (cofactor supply, modulators).

Transcriptional regulation of FMO involving receptors, ligand-binding and interaction

with DNA have not been widely studied as possible other forms of regulation.

Previously identified regulatory elements were found at the rabbit FMO1 promoter such

as the hepatic nuclear factor, homeodomain-containing factor (HNF1α), the hepatic

nuclear factor, orphan nuclear receptor (HNF4α) and the Yin Yang-1 (YY1) sites.

HNF1α and HNF4α were demonstrated to enhance FMO1 promoter activity, whereas

YY1 suppressed the ability of the upstream domains to enhance transcription.76 These

regulatory elements are conserved between rabbit and human FMO1. In rainbow fish,

hypersalinity induces FMO. Its regulatory pathway is not clear, but it was proposed that

cortisol directly regulate FMO. A glucocorticoid response element has been identified

in the promoter region of mammalian FMO1.77,78

A posttranslational regulatory effect could be the N-glycosylation of FMO1

selectively at Asn120.40 An NO modification of FMO3 activity in rat liver was also

reported. FMO3 activity and mRNA level were suppressed by NO, whereas dithiotreitol

and ascorbate were shown to restore FMO3 activity79.

A number of other mechanisms to regulate CYP including the enchancement of

mRNA stability, enzyme phosphorylation and protein-protein interactions either do not

occur for FMO or have not been reported.12

Dietary control

Rats switched from commercial chow diet to chemically defined diet (total parenteral

nutrition or synthetic aminoacid diet), exhibited a marked reduction in liver FMO

activity. These results suggested the existence of plant-derived constituents capable of

FMO induction.80 I3C and its gastric acid condensation products cause a marked

reduction in FMO activity.15 Total parenteral nutrition plus choline in rats caused a 3-

33

fold increase in hepatic microsomal FMO along with CYP2E1 induction.81 The FMO

activity was reduced by ca 30 % in starved female mice.82 Guinea pigs maintained on an

ascorbic acid free diet exhibited a significant decrease (45 %) in hepatic N-oxidation of

dimethylaniline.83

Developmental control

The most striking example for developmental control is the selective and antiparallel

expression of FMO1 and FMO3 in fetal and adult human liver. FMO1 is the major form

in human fetal liver, while FMO3 becomes the dominant form in adult liver. Human

FMO3 is not expressed in fetal liver, but it is induced after birth (0-8 years of age). The

signal for termination of FMO1 expression is related to the parturition and not to

gestation age.9 Other age-related changes in FMOs of human brain were shown only in

case of FMO1. It was also downregulated shortly after birth.40

In mice, FMO3 is presented 2 weeks after birth in both gender and remains until

puberty. Then, testosterone represses FMO3 enzyme production in males, while the

production does not change in females.38

Hormonal control

In rodents, FMO is regulated by steroids (sex steroids and glucocorticoids) and growth

hormone.84 Rat liver appears to be subject to positive regulation by testosterone and

negative control by estradiol.85 The expression of hepatic FMO1 seems to be repressed

by testosterone in male mice.38,85 Estradiol plays a role in the determination of FMO

activity in rat lung and kidney (FMO2 and FMO1).40 Cortisol regulates hepatic FMO

activity in female mice82 and also in rainbow fish77. FMO2 mRNA and protein

expression in pregnant rabbit correlates with progesterone and corticosterone. Up to a

20-fold variation of FMO activity has been observed in the corpora lutea of pigs during

estorus. Gestation increases FMO2 activity in rabbit86, mouse placenta87 and pig corpora

lutea88. The activity was also affected by the circadian rhythm, as it was lower ca 40 %

in the morning than in the afternoon in female mice.82

These observations have not yet been reported for humans. However, transient

trimethylaminuria was observed in females during menstruation.34

34

1.6.3 Pathophysiological status: diabetes

Diabetics are unable to absorbe glucose in certain cells because of insulopenia or insulin

resistance, leading to sugar accumulation in their blood.

Two classes of diabetes mellitus, type I and type II, with type II comprising over

80 % of the clinical cases, are known. Type I diabetes (also called juvenile or insulin-

dependent diabetes mellitus – IDDM) generally develops when patients are in the

adolescence and is characterized by the destruction of the β-cells, responsible for the

production of insulin, in the islet of Langerhans. Bacterial infection may set off an

immunological reaction in susceptibile people, which may initialize this form. Type II

diabetes (noninsulin dependent diabetes mellitus – NIDDM) generally develops later in

life and is caused by insulin resistance. Obesity and family history are prime risk factors

for the development of this form. Type I diabetes is treated with insulin

supplementation, while type II can often be controlled with diet, exercise and oral

hypoglycaemic agents available.89, ,90 91

CYP and FMO are the two major microsomal monooxygenase enzyme families

responsible for drug metabolism. Although many aspects of pathophysiological

consequences of diabetes mellitus have been scrutinized and its effect on drug

metabolism has been widely studied, there are only few reports regarding FMO. It

should also be noted that the effect of diabetes on various drug metabolising enzyme

systems depends on the type of the disease.

My thesis deals with the study of alteration of microsomal metabolic enzymes in

STZ-induced diabetic rats with type I form, therefore in the next part a brief survey will

be given on diabetes mellitus with special regard to pharmacokinetic and metabolic

aspects.

The effect of diabetes on pharmacokinetic parameters

For most drugs oral absorption is unlikely to be affected by diabetes, although a delay in

the absorption of tolazamide and a decrease in the absorption of ampicillin have been

reported. Subcutanous absorption of insulin is more rapid in diabetic patients, whereas

the intramuscular absorption of several drugs is slower.

The protein binding of a number of drugs in the blood is reduced in diabetes,

which may be due to the glycosylation of plasma proteins or displacement by plasma

35

free fatty acids, the level of which is increased in diabetic patients. Plasma

concentration of albumin and α1-acid glycoprotein does not appear to be changed by the

disease. The distribution of drugs with little or no protein binding in blood is generally

not altered, although the volume of distribution of antipyrine is reduced by 20 % in

IDDM.

Dainith in 1976 suggested that the clearance of antipyrine and other drugs could

be higher in diabetics92 (Table 5.). In addition, an elevated clearance of theophyllin was

found in type I93, but unchanged in type II diabetes94. The presence of fatty liver in

NIDDM may contribute to a reduced hepatic clearance, whereas decreased protein

binding in blood may cause an increased clearance.

The effect of diabetes on hepatic blood flow in humans appears to be unknown.

In diabetic adults the renal clearance of drugs either is comparable with that

found in nondiabetic individuals or reduced.95

Table 5. Pharmacokinetics of antipyrine in diabetics

Type of

diabetes

Total body

clearance

Volume of

distribution

Half-life

t1/2

I ↑ ↑ unchanged

II Unchanged ↑ ↑

I ↑ ↑ ↓

II Unchanged ↑ ↑

Antypirine was used as a probe drug for evaluating hepatic metabolic function. It is

metabolised by multiple CYP isoforms. 96,97

Drug metabolism in diabetes

The first clinical observation of altered drug metabolism in diabetic conditions was

reported in 1974, when Dajani found a slower conversion of phenacetin to paracetamol

in uncontrolled diabetic patients.98 Sotaniemi and co-workers showed in an

investigation using antipyrine showed that the drug metabolising characteristics of

diabetic patients significantly differ from those of non-diabetics and were dependent on

a number of factors such as age, gender, type of diabetes, effectiveness of antidiabetic

36

therapy and hepatic status, especially liver histology.99 Even in well-controlled diabetes,

drug metabolism may be altered, and the nature and extent of the changes are dependent

on the type of the disease.100 In human type II diabetes the expression of hepatic FMO5

isoform is markedly down-regulated.101 Alterations of FMO expression and activity in

diabetic patients could be of significance in drug therapy as a number of therapeutic and

recreational drugs are substrates for FMO to a certain extent.9

To elucidate the effect of diabetes on drug metabolism, studies on diabetic

animals were performed. The two primary ways of developing diabetic animals are their

treatment with streptozotocin (STZ) or alloxan in order to destroy pancreatic β-cells

responsible for insulin production. The mechanism of diabetogenesis regarding the two

agents is different. STZ at a dosage of 40-60 mg/kg administered by i.v. or >40-60

mg/kg i.p. causes IDDM while a dosage of 100 mg/kg on the day of birth causes

NIDDM. The latter form can also be modeled with congenital diabetic (Ob/Ob)

animals. The main reason for STZ induced β-cell death in the pancreas is DNA

alkylation.102

Alloxan at a dosage of 65 mg/kg administered by i.v. or at 150 mg/kg i.p. causes

diabetes in rats. 103 The toxic action of alloxan on pancreatic cells is a sum of several

processes such as oxidation of essential –SH groups, inhibition of glucokinase,

generation of free radicals and disturbances in intracellular calcium homeostasis. 104

STZ has a higher specificity as a β-cytotoxic agent in rat and is therefore reputed to be

less toxic as compared to alloxan.105

Animal studies on experimental type I diabetes revealed that diabetic state

modified the CYP-mediated hepatic metabolism of certain drugs. 106 It was reported that

codein, chlorpromazine, hexobarbital and phenylbutazone were metabolized more

slowly in alloxan-diabetic male rats than in control animals.106,107 The elimination rate

of chlorpropramide fell rapidly in diabetic animals.108 The microsomal metabolism of

aniline was reported either increased109 or unchanged110. In addition, the FMO-mediated

imipramin N-oxidation in the liver increased 1.5- to 2-fold in comparison to control

mice both in congenital and STZ-induced diabetes.111

Focusing on streptozotocin-induced diabetic rats, changes in drug metabolizing

capacity of hepatic monooxygenases have been reported. There are several conflicting

reports on the changes in hepatic CYP content and isozyme activities. Both, increase112

37

and absence of change113 of total cytochrome P450 content in diabetes were reported.

An elevated cytochrome b5 level was detected in diabetic rats in the latter study. In case

of CYP2B, CYP3A and CYP2E1 both induction112, ,114 115 and repression have been

reported110

14

2

, ,112116 . The induction of CYP1A activity was confirmed by independent

studies.113,1 The most likely explanation for these discrepancies is that the metabolic

status of experimental animals -just as it was mentioned above for humans - is strongly

influenced by the varying severity of induced diabetes. In addition, strain, sex and age

may also introduce some variations.117 In diabetic rats, Rouer and her co-workers

detected two-fold higher FMO activity than in normal rats, whereas they could not show

any changes in FMO protein expression. However, a slight structural difference of the

enzymes was indicated by the analysis of their tryptic peptide profiles.118 In the study of

Wang the hepatic FMO1 protein expression and activity were induced 2.5-fold and 1.8-

fold in diabetic rats compared to non-diabetic rats.119 Based on animal studies, it was

suggested that in diabetes the higher FMO activity might be due to the presence of an

endogenous factor120, a post-translational enzymatic activation or changes in

conformation of the enzyme in the membrane caused by changes of its lipid

composition.117,121 Rouer proposed that the activity of the flavin monooxygenase

appears to be controlled by a glycoregulatory hormone.122

The effect of insulin supplementation was studied in diabetic rats. The change of

the hepatic metabolic rate of aminopyrine on insulin treatment was reversible11

whereas that of aniline and testosterone were not115. Insulin treatment itself caused a

significant change only in the activity of ethoxyresorufin O-deethylase.123 Yamazoe

suggested that insulin might exert an indirect effect on CYPs through the normalization

of growth-hormon-mediated process(es) in diabetic rats.114

Enhanced CYP2E1 protein expression and suppressed CYP2B and CYP3A

expressions were seen in the absence of insulin in primary rat hepatocyte culture.124

Moreover, Woodcroft had demonstrated that insulin suppresses CYP2E1 transcription.

Insulin as a regulator of flavin-containing monooxygenase in experimental diabetes

has not yet been demonstrated.

38

2 RESEARCH OBJECTIVES

Our main goal was to study if insulin has a role in FMO regulation. For this purpose we

induced experimental diabetes using streptozotocin as diabetogenic agent. Rat was

selected as a model animal. Rats were divided into 5 groups: 1. control, 2. STZ-induced

diabetic (lower dosage), 3. STZ-induced diabetic (higher dosage), 4. insulin treated

diabetic, 5. insulin treated non-diabetic rats. Changes in FMO function was determined

at two levels: enzymatic activity and gene expression. FMO activity was measured

using BZY as an FMO specific substrate and the BZY N-oxide produced was analysed

and quantified by an HPLC method applying UV detection developed by us. The gene

expression of FMO1 and FMO3 were followed by quantitative Real-time PCR. Along

that, the changes in abundance of cytochromes (cytochrome P450 and cytochrome b5)

and the enzymatic activities of hepatic CYP1A, CYP2B, CYP2E1 and CYP3A were

characterized in order to support and complete the results in our experimental model

system. The objectives were as follows:

1. Does insulin have a role in FMO regulation?

2. Does insulin treatment in STZ-induced diabetic rats restore FMO function?

3. Does the severity of diabetes show a significant correlation with FMO function?

4. Does insulin itself regulate FMO in non-diabetic rats?

5. Does insulin regulate the function of FMO at enzymatic and/or gene expression

level?

6. Are both FMO isoforms equally sensitive to insulin-deficiency?

7. Does FMO activity correlate with the activity or abundance of any other

cytochrome enzymes altered in experimental diabetes?

39

3 MATERIALS AND METHODS

3.1 BIOCHEMICALS

All biochemicals used were of the highest purity available from commercial sources.

Streptozotocin, sodium dithionite, 6β-hydroxytestosterone, aminopyrine, glucose-6-

phosphate, glucose-6-phosphate dehydrogenase, ethoxyresorufin and benzydamine were

obtained from Sigma-Aldrich (St. Louis, MO), MpTS and its sulfoxide was from

Sigma-Aldrich Chemie GmbH (Steinheim, Germany). Benzydamine N-oxide

metabolite was synthetized in the Synthetic Chemistry Research Laboratory II of

Gedeon Richter Ltd (Budapest, Hungary) by an established method.125 It was identified

and characterized by comparison of melting point and TLC, NMR and MS data to that

of the authentic samples. Tolperisone hydrochloride and metabolite standards were

synthesized in Gedeon Richter Ltd (Budapest, Hungary). Resorufin and 4-nitrocatechol

were from Aldrich Chemical Co. (Milwauuke, WI). Testosterone was from Fluka

(Buchs, Switzerland). Para-nitrophenol, sodium pyrophosphate, MgCl2, NADPH,

perchloric acid, sodium bicarbonate, phosphoric acid and acetic acid were obtained

from Reanal (Budapest, Hungary). Insulin (Ultratard) was purchased from Novo

Nordisk (Novo Allé, Denmark). Human liver microsomes for studying deprenyl

metabolism were obtained from Xenotech LLC (Kansas City, KS, USA), while for

tolperisone studies from In Vitro Technologies (Baltimore MD). Microsomes

containing recombinant human FMO1 and FMO3 from baculovirus-insect cell

(supersomes) were purchased from BD Gentest (Woburn, MA, USA). Hydrochloride

salts of S- and R-methamphetamine, S- and R-amphetamine, S- and R-deprenyl, S- and

R-methamphetamine-hydroxylamine, S- and R-amphetamine-hydroxylamine, 1R,NR-

and 1S,NS-deprenyl N-oxide, and the diastereomeric mixture (1:1) of 1R,NR and

1R,NS- and 1S,NS- and 1S,NR-deprenyl N-oxide, and S-para-fluorodesmethyldeprenyl

(internal standard) were kindly provided by Chinoin Pharmaceutical and Chemical

Works, a member of the Sanofi-Aventis Group (Budapest, Hungary). CE-grade 2-

(hydroxy-propyl)-β-cyclodextrin was supplied by Beckman Coulter (Fullerton, CA,

USA). Hydroxypropylmethylcellulose was obtained from Sigma (St. Louis, MO).

Discovery C18 Lt SPE cartridges (Supelco, Bellefonte, PA, USA) were used for solid-

40

phase extraction of the samples. Chromatographic solvents were of gradient grade and

were purchased from Merck (Darmstadt, Germany).

3.2 ANIMALS AND INDUCTION OF DIABETES

The study was carried out in accordance with the Declaration of Helsinki. Male

Sprague-Dawley rats (160-170 g) were purchased from Harlan (Holland). Rats were

maintained at 20-25 °C on a 12-h light/12-h dark cycle, with access to water and food

ad libitium. Diabetes was induced by a single intraperitoneal injection of buffered

solution (0.1 mol/l citrate, pH 6.0) of streptozotocin at doses of 50 and 70 mg/kg. The

animals were considered diabetic if their blood glucose concentrations elevated over

350 mg/dl on the 4th day following streptozotocin treatment. The animals were

separated in 5 groups: control (n=8); diabetic rats treated with streptozotocin at

50 mg/kg dose (n=9), D50; diabetic rats treated with streptozotocin at 70 mg/kg dose

(n=8), D70; diabetic rats treated with streptozotocin at 70 mg/kg dose and from the 5th

day insulin treatment was initiated with dose gradually increasing (on day 1: 20,

day 6: 40, day 7: 50 and from the day 8: 2x75 IU/kg/day) (n=8), ID70; and non-diabetic

rats with insulin treatment daily at a dosage of 50 IU/kg for avoiding hypoglycaemia

(n=5), IND. During insulin treatment blood glucose level was checked twice daily using

Glucotrend kit (Mannheim, Germany). Animals were sacrificed by cervical dislocation

2 weeks after the streptozotocin treatment in a short interval so as to avoid circadian

variation.

3.3 PREPARATION OF LIVER MICROSOMES

After isolating the liver it was homogenised in 1.15 % KCl containing Tris-HCl buffer

(0.1 M, pH 7.4) with the volume/mass ratio of 2:1 buffer (V, ml) and liver (m, g). The

buffer also contained 0.2 mM ethylenediamine-tetraacetic acid (EDTA), 0.1 mM

dithiotreitol (DTT), 0.1 M tris hydroxymethyl aminomethane (Tris), 1.15 % of KCl and

10 % of glycerine. The pH was adjusted with HCl to 7.4.

Individual rat liver microsomes (RLM) were prepared by differential

centrifugation (the homogenate was centrifuged for 15 min at 9000 x g, the supernatant

41

was ultracentrifugated for 60 min at 105000 x g). The microsomal pellet obtained was

resuspended in the same buffer.

The microsomal protein content was measured with alkaline folin phenol reagent

with bovine serum albumin as standard using the method described by Lowry.126

Therefore, 2 ml mixture of three reagents (A: 10 ml 0.2 N NaOH + 4 % Na2CO3, B: 0.2

ml 2 % KNa- tartrate, C: 0.2 ml 1 % CuSO4) were added to an equal volume of the

microsomal suspension (0.25 %). Ten minutes later 0.2 ml of Folin-Ciocalteu phenol

reagent was added. After 30 minutes of exposure the absorbance of violet Cu-complex

in the samples was determined at 750 nm by spectrophotometer (Beckman DU-30).

3.4 DETERMINATION OF CYTOCHROME CONTENT

The content of cytochromes (i.e. cytochrome P450 and cytochrome b5) was essentially

determined by the method described by Greim (1970).127 Briefly, 1.2 ml of microsomal

suspension was bubbled by carbon monoxide for 15-20 seconds (20-30 bubbles) and

divided into two aliquots (i.e. sample and reference cell). The aliquot in the sample cell

was reduced by a few crystals of solid dithionite and the differential spectra were

recorded. Cytochrome P450 concentration (nmol/mg protein) was calculated from the

absorbance difference measured at 450 and 490 nm using ε=91 mM-1cm-1, whereas the

concentration of cytochrome b5 was calculated from the same spectra from the

difference of absorbences measured at 409 and 424 nm by the use of ε=161 mM-1cm-1

(Hitachi U-3300).

3.5 ENZYMATIC ASSAYS

All incubations were carried out in incubation mixtures containing 6.25 mM sodium

pyrophosphate, 5 mM MgCl2, 5 mM glucose-6-phosphate, 1 U/ml glucose-6-phosphate

dehydrogenase and 0.25 to 1 mg/ml RLM in a final volume of 0.5 or 2 ml of 0.1 M

Tris-HCl buffer, pH 7.4. Incubation time and microsomal protein concentrations were

within the linear range of metabolite formation.

42

3.5.1 FMO index reaction: Benzydamine N-oxygenation

BZY/BZY N-oxide transformation was used as an FMO specific index reaction. BZY

N-oxide was measured by a modified method of Kawaji50 (see in Results).

NN

O

N

NN

O

N

O

FMO1 / FMO3

Fig. 8. FMO-mediated BZY N-oxide formation

The metabolic profile of BZY transformation in rat liver microsomes was also

studied. The analytical measurement was performed on a Thermo Finnigan Surveyor

HPLC system equipped with UV and MS detector. A Superspher 60 Rp-select B,

4x125 mm, (5 µm) column was operated at 0.8 ml/min flow rate, maintained at 30 °C.

The metabolites of BZY were determined using a mobil phase of a mixture of methanol

and 0.1 M ammonium-acetate in the presence of 0.1 % AcOH, using a gradient of 20:80

(A) to 80:20 (B); gradient steps: A of 60 (2 min), 10 (6-13 min), and 60 % (13.1-15

min). The UV signal was detected at 306 nm. 10 µl supernatant was injected onto

HPLC. The MS was set up for positive ionization (ESI) with an ionising potential of 3

kV; the temperature of the capillary was 220 ºC.

3.5.2 CYP1A index reaction: Ethoxyresorufin O-deethylation

CYP1A1/2 specific activity was measured by a modified method of Burke.128 In a final

volume of 2 ml the substrate concentration was 100 µM, the microsomal protein

concentration 1 mg/ml. Prior to the initiation of the reaction by the addition of 7-

ethoxyresorufin a 5 minutes long preincubation was taken along. The incubation was

carried out for 15 minutes in shaking water bath at 37 °C. After the indicated time the

43

reaction was stopped with 1 ml of ice-cold methanol. Then the samples were

centrifuged for 30 minutes at 3000 x g, at 4 °C. The supernatant was then analysed by

fluorometer (Hitachi F-4000) as the metabolite, resorufin, is fluorescent (λex=535 nm

and λem=585 nm). The metabolite concentration was determined based on the

calibration of resorufin at 125, 250, 500, 1000, 1500, 2000 and 2500 nM.

3.5.3 CYP2B/3A index reaction: Aminopyrine N-demethylation

Aminopyrine N-demethylation was determined by measuring formaldehyde formation

via the method of Nash.129 The substrate concentration was 4 mM, the microsomal

protein concentration used was 1 mg/ml in a final volume of 2 ml. The reaction was

initiated by the addition of NADPH. The incubation was carried out for 30 minutes in

shaking water bath at 37 °C. After the indicated time the reaction was stopped with

500 µl of 20 % TCA. Then the samples were centrifuged for 20 minutes at 3000 x g, at

4 °C. In order to develope the colour reaction 2 ml Nash reagent was added to 1.5 ml of

supernatant and left them in waterbath at 37 °C, without shaking. After 30 minutes, the

samples were spectrophotometrically quantified at 410 nm (Beckman DU-30). The

produced formaldehyde concentration was determined based on the calibration of

HCHO at 20, 40, 80, 160 and 400 µM.

3.5.4 CYP2E1 index reaction: p-Nitrophenol-hydroxylation

CYP2E1 specific activity was measured by a modified method of Reinke and Moyer.130

The substrate concentration was 2 mM, the microsomal protein concentration used was

1 mg/ml in a final volume of 2 ml. The reaction was initiated by the addition of

NADPH. The incubation was carried out for 15 minutes in a shaking water bath at

37 °C. After the indicated time the reaction was stopped with 500 µl of 20 % TCA.

Then, the samples were centrifuged for 20 minutes at 3000 x g, at 4 °C. In order to

develope the colour reaction 1 ml 2 M NaOH reagent was added to 1.5 ml of

supernatant. The samples were quantified at 505 nm wavelength (Beckman DU-30).

The produced p-nitrocatechol (PNC) concentration was determined based on the

calibration of PNC at 2.5, 5, 10, 20, 30, 40, 50, 60 and 75 µM.

44

3.5.5 CYP3A index reaction: Testosterone-6β-hydroxylation

CYP3A specific activity was measured by a modified method of Anderson.131 The

substrate concentration was 500 µM, the microsomal protein concentration used was

0.25 mg/ml in a final volume of 500 µl. The reaction was initiated by the addition of

NADPH. The incubation was carried out for 5 minutes at 37°C using a shaking water

bath. After the indicated time the reaction was stopped with 500 µl of ice-cold

methanol. Samples were placed into a refrigerator (-20 °C) for 10 minutes. Then the

samples were centrifuged for 10 minutes at 10000 x g, at 4 °C. The supernatant was

injected onto HPLC. The analytical measurement was performed on a Merck-Hitachi

LaChrom HPLC system equipped with UV detector. Purospher C18e 125 x 4 mm (5

µm) column (Merck) operated at 0.8 ml/min flow rate, maintained at 40 °C. 6β-

hydroxytestosterone was determined using mobil phases of methanol/0.1 M

ammonium-acetate, 45:55 (A) and methanol (B) with a gradient (A of 100 (4 min), 40

(14-17 min), and 100 % (20-26 min)). UV detection was at 254 nm.

The metabolite concentration was determined based on the calibration of 6β-

hydroxytestosterone at 1, 5 and 25 µM.

3.6 DETERMINATION OF THE MRNA OF FMO1 AND FMO3 ISOFORMS

3.6.1 Isolation of total RNA

The isolation was carried out using RNeasy Mini Kit (Qiagen) by following the given

protocol. Hence, 30-30 mg of liver sample of each examined animal was placed into

500 µl aliquotes of RNA stabilization reagent. The tissue was disrupted and

homogenized in Potter with lysing buffer, containing highly denaturating guanidine

isothiocyanate. The tissue lysate was centrifuged at 10000 x g for 3 minutes. The

supernatant was transferred to a new microcentrifuge tube and 500 ul of 70 % ethanol

was added. Then 700 µl of the sample was applied to an RNeasy mini column (silica-gel

membrane) and DNA and protein were eluted 3 times with washing buffer. The mRNA

was eluted by 50 µl of RNase-free water. This procedure provides enrichment for

mRNA.

45

3.6.2 Measurement of concentration and integrity of RNA

The concentration and integrity of purified RNA was checked by Agilent 2100

bioanalyser using RNA 6000 Nano Assay kit. RNA quality is characterized by a

number called RNA integrity number (RIN), given in a scale of 0-10.

3.6.3 Synthesis of cDNA

Full-length first strand cDNA was generated from an RNA template using Ready–To-

Go You-Prime First-Strand Beads (Amersham Pharmacia Biotech) with addition of 2

µg of denaturated total RNA sample and random primers (Promega). The reaction was

carried out at 37 ˚C for 60 minutes.

3.6.4 Gene expression analysis

The quantification of mRNA levels was performed using Applied Biosystem TaqMan®

Gene Expression Assays. One-tube TaqMan Real-time PCR was ran using Taqman

Universal PCR master Mix (Applied Biosystems) and standard conditions in 25 µl

reactions on the ABI PRISM® 7000 Sequence Detection System (Applied Biosystems).

Specific amplifications of cDNA was performed using TaqMan® Gene Expression

Assays consist PCR primers and FAM™ dye-labeled TaqMan® MGB probe. Taqman

assays for rFMO1 (Rn00562945_m1) rFMO3 (Rn00584825_m1) and huGAPDH

(Applied Biosystems) were used. GAPDH was used as an endogenous control to

normalize for differences in amount of total RNA added to the reaction. Treshold cycle

(Ct) was determined for GAPDH, FMO1 and FMO3. The measurements were done two

times in triplicates. The relative amount of FMO1 and FMO3 mRNA was determined

by comparative Ct method as the following:

CtGAPDH = average for GAPDH of the sample in question

CtFMO= individual value for FMO of the sample in question

∆Ct=CtFMO-CtGAPDH

∆∆CtFMOtreated=∆CtFMOtreated-average(∆CtFMO control)

Fold-induction FMO treated=2-∆∆CtFMOtreated

46

3.6.5 Data analysis

Statistical analysis was performed by Statistica 6.0 (StatSoft, Inc.). Normality test

(Shapiro-Wilk’s test) and homogenity test (Bartlett’s test) were run for all datasets in

order to determine whether they have Gaussian distribution and homogenity in their

variances. In case of parametric distribution one-way analysis of variance (ANOVA)

was used which was followed by Tukey multicomparison test if any significant change

was shown. This procedure was followed for all enzyme activities and content datasets

except for FMO activity and mRNA values. The latter ones had non-parametric

distribution (despite transformations), therefore for their analyses Kruskal-Wallis

multicomparison test was used. Mean, S.D. and correlation coefficient were calculated

by GraphPad Prism 2.01 (GraphPad Software, Inc.). The confidence interval was

chosen to 95 %.

3.7 FMO-MEDIATED BIOTRANSFORMATION OF SOME CNS DRUGS

3.7.1 Tolperisone

All incubations were carried out in incubation mixture described in chapter 3.5.

Tolperisone concentration used was 5 µM (inhibition experiment) or 0-1500 µM

(enzyme kinetic measurement). In incubations with human liver microsomes (HLM),

the mixture contained 1 mg/ml of pooled microsomal protein and the reaction was

initiated by the addition of NADPH and conducted using a shaking water bath at 37 ˚C

for 20 minutes. Tolperisone consumption and major metabolite formation was measured

by HPLC. In incubations containing insect cell-expressed recombinant human FMO3

the enzyme concentration was 1 nmol/ml. The reactions were initiated by the addition of

microsomes and incubated in water bath at 37 ˚C, without shaking, according to the

manufacturer’s recommendation. Microsomes prepared from wild-type baculovirus

infected cells were used as control. The reaction was stopped with equal volume of ice-

cold methanol, centrifuged for 10 min at 7500 x g, and 10 µl of supernatant was injected

into HPLC. Thiourea at 10 mM was used as a competitive inhibitor of FMO3 in an

inhibition study carried out with HLM. Tolperisone was used as an assumed inhibitor

and MpTS/MpTSO transformation as an FMO index reaction125 for determining Ki of

47

tolperisone using insect cell-expressed recombinant human FMO3 enzyme. Incubations

were performed for 20 minutes. MpTSO was analysed with HPLC.

Analytical measurements were carried out using a Merck-Hitachi LaChrome HPLC

system equipped with UV detector. Discovery C18 150x4.6 mm (5µm) column with

Supelguard 20x4 mm (5 µm) column (Supelco, Bellefonte, PA) was maintained at 40 ˚C

for all assays. Tolperisone and hydroxymethyl-tolperisone was monitored at 256 and

251 nm, respectively. The system was operated at 0.8 ml/min flow rate, under isocratic

conditions. The mobile phase contained 47 % methanol in 0.1 M ammonium-acetate.

MpTSO was monitored at 233 nm using methanol/0.1 M ammonium-acetate, 34:46 (A)

and methanol (B) as mobile phases with a gradient grade of (A) 100 (5 min), 35 (12

min) and 100 % (16-22 min).

3.7.2 Deprenyl, amphetamine, methamphetamine enantiomers

All incubations were carried out in the incubation mixture described previously in

chapter 3.5. The concentration of substrates was 1 mM. In incubations with human liver

microsomes, the mixture contained 1 mg/ml microsomal protein and the reaction was

initiated by addition of NADPH (0.5 mM final concentration), and carried out in

shaking water bath at 37 ˚C for 30 min. For the incubations containing insect cell-

expressed recombinant human FMOs the enzyme concentration was 100 nM (FMO1

and FMO3). The reactions were initiated by the addition of supersomes and incubated

in water bath at 37 ˚C, without shaking, in accordance with the manufacturer’s

recommendation. Microsomes prepared from wild-type baculovirus infected cells were

used as control. The reaction was stopped with equal volume of 0.4 M cold perchloric

acid and centrifuged for 10 min at 10000 x g. After solid-phase extraction, the residue

redissolved in distilled water was analysed by capillary electrophoresis (CE). All

experiments were performed with a P/ACE MDQ CE system controlled by 32 Karat

software, Version 5.0 (Beckman Coulter). The instrument was equipped with UV

detector, set at 200 nm. Separations were carried out in fused-silica capillaries, 75 µm

ID, 365 µm OD, 50.2 cm total length, 40 cm to the detector (Polymicro Technologies,

Phoenix, AZ, USA). Samples were introduced into the capillary by pressure (5s, 0.5 psi)

with the sample volume corresponding to about 1.25 % of the capillary volume. The

48

separations were carried out applying a constant voltage of 25 kV, and the temperature

was maintained at 15 ˚C.

49

4 RESULTS

4.1 DRUG METABOLIZING MICROSOMAL ENZYMES IN DIABETIC RATS

4.1.1 Changes of physical and biochemical parameters in diabetic rats

Streptozotocin treated rats became diabetic (25 animals out of 27) with symptoms of

polydipsia, polyuria, hyperglycaemia and decreased rate of body weight gain. The

measured physiological parameters of the animals are summarized in Table 6. Control

and insulin treated non-diabetic (IND) rats consistently had blood glucose levels of

approx. 100 mg/dl and 110 mg/dl respectively, while rats receiving streptozotocin at

50 and 70 mg/kg doses (in order to reach a broader interval of blood glucose level for

correlation studies) displayed about a five-fold increase in blood glucose levels 520 and

540 mg/dl, respectively. Diabetic rats had a dose-dependent reduction of body weight

and wet liver weight compared to controls, at the time of sacrifice. Insulin treatment of

diabetic animals only partially compensated the rise of blood glucose levels

(390 mg/dl), but fully the changes of body weight and wet liver weight, which resulted

in an enhanced relative liver weight. The body weight and the wet liver weight of IND

animals were reduced in comparison to control animals.

50

Anim

als w

ere

trea

ted

as d

escr

ibed

in M

ater

ials

and

Met

hods

. Mea

n ±

S.D

. wer

e ca

lcul

ated

(n=

5-9)

. *p ≤

0.05

, **

p ≤

0.01

, ***

p ≤

0.00

1 vs

. con

trol

val

ues,

†p≤

0.05

val

ues o

f D50

vs.

D70

, ‡p ≤

0.05

, ‡‡‡

p ≤

0.00

1 va

lues

of D

70 v

s. ID

70.

Tab

le 6

. Eff

ects

of s

trep

tozo

toci

n-in

duce

d di

abet

es a

nd in

sulin

trea

tmen

t of d

iabe

tic a

nd n

on-d

iabe

tic r

ats

on v

ario

us p

hysi

cal a

nd b

ioch

emic

al c

hara

cter

istic

s

51

4.1.2 The function of FMOs

Benzydamine N-oxide is an FMO-specific metabolite in rat liver

The activity of FMO was measured using an FMO specific BZY/BZY N-oxide marker

reaction. The FMO specific formation of BZY N-oxide with rat liver microsomes was

confirmed in our metabolic profile study. BZY N-oxide was observed to be the major

metabolite, whereas desmethyl-BZY was produced in a lower concentration. There were

two additional minor metabolites, which were suggested to be monohydroxylated

derivates of BZY (Fig. 9.).

D:\Xcalibur\data\2881 11/08/2005 01:47:16 PM benzidamin 500uM 60

RT: 2.01 - 14.03

3 4 5 6 7 8 9 10 11 12 13 14Time (min)

5

10

15

20

25

30

35

40

45

50

55

60

65

70

75

80

85

90

95

100

Rel

ativ

eA

bsor

banc

e

RT: 7.20MA: 3518785

RT: 8.31MA: 1234771

RT: 6.96MA: 266243

RT: 5.40MA: 53941

RT: 6.28MA: 47667

NL:5.58E5Channel AUV 2881

NN

O

NH CH3

M=295

NN

O

NCH3CH3

O

M=325

NN

O

NCH3

CH3

M=309

benzydamine

benzydamine- -oxideN

N-demethyl-benzydamine

benzydamine + OM= 325

Fig. 9. Metabolites of BZY incubated with rat liver microsomes

Metabolites were separated by HPLC, detected by UV, identified using MS. The molecular weights of different metabolites are indicated in the chromatogram.

52

Development of a new HPLC-UV method for the determination of benzydamine N-oxide

Based on the study of substrate-, protein- and time-dependency of BZY/BZY N-oxide

transformation, the parameters of incubation were all within the linear range. The

substrate concentration was 500 µM, the microsomal protein concentration used was

0.25 mg/ml in a final volume of 500 µl. The reaction was initiated by the addition of

NADPH. The incubation was carried out for 5 minutes using a shaking water bath at

37 °C. After the indicated time the reaction was stopped with 500 µl of ice-cold

methanol. Samples were placed into a –20 °C fridge for 10 minutes. Then, the samples

were centrifuged for 10 minutes at 10000 x g, at 4 °C followed by injection of 10 µl

supernatant onto HPLC. The analytical measurement was performed on a Merck-

Hitachi LaChrom HPLC system equipped with UV detector. Purospher C18e 125 x 4

mm (5 µm) column (Merck) operated at 0.8 ml/min flow rate, maintained at 30 °C.

Benzydamine N-oxide was monitored at 306 nm. The mobile phase consisted of 58 %

methanol in 0.1 M ammonium-acetate and isocratic elution was applied. Under these

chromatographic conditions, BZY and BZY N-oxide eluted at 9.5 and 6.8 min,

respectively. The metabolite concentration was determined based on the calibration of

BZY N-oxide at 2, 5 and 25 µM. All measurements were carried out in triplicates.

Specific enzymatic activity of FMO

FMO activity elevated in D50 and D70 rats (p=0.0008 and p=0.0001, respectively). It

increased in a streptozotocin dose-dependent manner: 50 mg/kg and 70 mg/kg doses

caused 39 % and 73 % increase, respectively. In insulin treated diabetic (streptozotocin

70 mg/kg) (ID70) animals, the FMO activity was restored to the control level. Insulin

had no effect on the FMO activity of IND animals. (Fig. 10.).

53

Control D50 D70 ID70 IND0

2000

4000

6000

8000

10000

***

***†

‡‡

FMO activity

FMO

spe

cific

act

ivity

(pm

ol*m

g-1*m

in-1

)

Fig. 10. Hepatic FMO activity in streptozotocin-induced diabetic rats with or

without insulin treatment and in non-diabetic insulin treated animals Incubations were carried out in triplicates. Mean ± S.D. were calculated (n=5-9).

***p ≤ 0.001 vs. control value, †p ≤ 0.05 values of D50 vs. D70, ‡‡p ≤ 0.01 values of D70 vs. ID70.

Gene expression of FMO1 and FMO3

The RNA purification and cDNA synthesis

Rneasy Mini Kit (Qiagen) was used for the isolation of the total RNA. The major

features for a high quality total RNA run were two ribosomal peaks (18S and 28S)

accompanied by a low level of degradation products, which in higher concentration

would have caused an increased baseline. All RNAs prepared from the samples

analysed were of high quality. As an example, an electropherogram of a typical sample

(RIN=10) is shown below (Fig. 11.). The cDNA was synthesized using 2 µg of total

RNA.

54

28S

18S

Fig. 11. The electropherogram of an RNA sample The two characteristic ribosomal peaks can be seen.

Gene expression of FMO1 isoform

In rats treated with 70 mg/kg of STZ the mRNA level of FMO1 increased by 84 %. The

gene expression was restored to the control level as a result of insulin treatment. Insulin

itself did not cause any changes in the FMO status (Fig. 12.).

FMO1 mRNA

Control D50 D70 ID70 IND0

1

2

3

*

‡‡‡

rela

tive

amou

nt o

f FM

O1

mR

NA

Fig. 12. Hepatic FMO1 mRNA level in streptozotocin-induced diabetic rats with or

without insulin treatment and in non-diabetic insulin treated animals Mean ± S.D. were calculated (n=5-9). *p ≤ 0.05 vs. control value, ‡‡‡p ≤ 0.00 1 values

of D70 vs. ID70.

55

Gene expression of FMO3 isoform

The FMO3 mRNA level of rats that received 50 mg/kg and 70 mg/kg doses of STZ

increased by 2- and 4-fold, respectively. As a result of insulin treatment the gene

expression was restored to the control level. Insulin itself did not cause any changes in

FMO3 status (Fig. 13.).

FMO3 mRNA

Control D50 D70 ID70 IND0

1

2

3

4

5

6

*

***

rela

tive

amou

nt o

f FM

O3

mR

NA

Fig. 13. Hepatic FMO3 mRNA level in streptozotocin-induced diabetic rats with or

without insulin treatment and in non-diabetic insulin treated animals Mean ± S.D. were calculated (n=5-9). *p ≤ 0.05, ***p ≤ 0.001 vs. control value,

‡p ≤ 0.05 values of D70 vs. ID70.

Changes of FMO function

The FMO specific enzyme activity, FMO1 and FMO3 mRNA levels are plotted in

Fig. 14. For the FMO activity measured all FMO isoforms present in rat liver are

responsible since the BZY N-oxide formation is only an FMO specific, but not FMO

isoform specific reaction. The mRNA level of FMO3 elevates more dynamically than

that of FMO1. Since FMO1 is the dominant isoform represented in rat liver, the

catalytic efficacy for both enzymes are very similar and the patterns of change of FMO

activity and FMO1 mRNA level match each other perfectly, it is suggested that the

FMO1 isozyme is responsible for the change of the total FMO activity in rats. The

FMO1 and FMO3 isoforms are regulated by insulin to a different extent.

56

57

FMO

1 m

RNA

Con

trol

D50

D70

ID70

IND

0123456

*

‡‡‡

relative amount of FMO1 mRNA

FMO

3 m

RN

A

Con

trol

D50

D70

ID70

IND

0123456

*

***

relative amount of FMO3 mRNA

FMO

act

ivity

Con

trol

D50

D70

ID70

IND

0123456

‡‡

† ***

***

relative FMO activity

FMO

1 m

RNA

Con

trol

D50

D70

ID70

IND

0123456

*

‡‡‡

relative amount of FMO1 mRNA

FMO

3 m

RN

A

Con

trol

D50

D70

ID70

IND

0123456

*

***

relative amount of FMO3 mRNA

FMO

act

ivity

Con

trol

D50

D70

ID70

IND

0123456

‡‡

† ***

***

relative FMO activity

Fig.

14.

Com

pari

son

of F

MO

act

ivity

, FM

O1

and

FMO

3 m

RN

A le

vels

Bo

th, t

he F

MO

act

ivity

and

FM

O m

RNA

leve

ls w

ere

incr

ease

d in

dia

betic

rats

whi

ch d

eclin

ed to

con

trol

leve

l on

insu

lin tr

eatm

ent.

Ther

e w

ere

no c

hang

es o

bser

ved

in F

MO

stat

us in

non

-dia

betic

rats

trea

ted

with

insu

lin p

er se

. *p ≤

0.05

, ***

p ≤

0.00

1 vs

. con

trol

va

lues

; †p ≤

0.05

val

ues o

f D50

vs.

D70

; ‡p ≤

0.05

, ‡‡p

≤ 0

.01,

‡‡‡

p ≤

0.00

1 va

lues

of D

70 v

s. ID

70.

4.1.3 The function of cytochromes: CYP and cytochrome b5

Total cytochrome P450 content and cytochrome b5 concentration

Cytochrome P450 content in ID70 animals decreased by 34 % and 36 % in comparison

to control and diabetic (streptozotocin, 70 mg/kg) (D70) rats, respectively. In contrast to

cytochrome P450, major changes were shown in cytochrome b5 concentration in

untreated diabetic rats. The cytochrome b5 concentration increased in diabetic rats in a

streptozotocin dose-dependent manner: 50 mg/kg and 70 mg/kg doses caused 50 % and

95 % increase, respectively. The cytochrome b5 concentration in ID70 animals was

restored to control level, while in IND animals it remained unchanged (Table 7.).

Cytochrome P450 isozyme activities

The ethoxyresorufin O-deethylase activity, used as CYP1A marker reaction, increased

in diabetic rats by 48 % and 67 %, and it was restored to control values by insulin

treatment (Table 7.). In IND animals this CYP1A-mediated activity did not change. The

aminopyrine N-demethylase activity, a nonspecific index reaction of CYP2B/CYP3A,

decreased in diabetic rats in a streptozotocin dose-dependent manner: 50 mg/kg and

70 mg/kg doses caused mild reduction of activity. Aminopyrine N-demethylase activity

in ID70 animals was not restored to control level and there was a mild and significant

decrease in IND animals. The p-nitrophenol hydroxylase activity, an index reaction of

CYP2E1, did not change substantially either in ID70 or D70 rats. In contrast, there was

a marked 73 % decrease in CYP2E1 activity in IND rats. The CYP3A-mediated

testosterone 6β-hydroxylase activity increased by 43 % only in diabetic (streptozotocin,

50 mg/kg) (D50) rats. There was no change in CYP3A activity in either of the insulin

treated groups.

58

Tab

le 7

. Hep

atic

cyt

ochr

ome

P450

con

tent

and

isoz

yme

activ

ities

in st

rept

ozot

ocin

-indu

ced

diab

etic

rat

s with

or

with

out i

nsul

in tr

eatm

ent a

nd in

non

-dia

betic

insu

lin tr

eate

d an

imal

s

Det

erm

inat

ions

wer

e do

ne a

s des

crib

ed in

Mat

eria

ls a

nd M

etho

ds. I

ncub

atio

ns w

ere

carr

ied

out i

n pa

ralle

l and

m

ean ±

S.D

. wer

e ca

lcul

ated

(n=

5-9)

. *p ≤

0.05

**p

≤ 0

.01,

***

p ≤

0.00

1. v

s. co

ntro

l val

ues,

††p ≤

0.01

val

ues o

f D

50 v

s. D

70, ‡

p ≤

0.05

, ‡‡p

≤ 0

.01,

‡‡‡

p ≤

0.00

1 va

lues

of D

70 v

s. ID

70.

59

4.1.4 Correlation analysis

The directions and extents of the changes in blood glucose level, FMO activity and

cytochrome activities or contents (Table 8.) motivated us to perform regression analysis

between these parameters.

Table 8. The changes in blood glucose concentration, FMO activity, cytochrome

contents and activities in streptozotocin-induced diabetic rats with or without

insulin treatment and in non-diabetic insulin treated animals

Treatment BG FMO Total CYP Cyt.b5 CYP1A CYP2B/

CYP3A CYP2E1 CYP3A

D50 ↑↑ ↑ UC ↑ ↑ ↓ UC ↑

D70 ↑↑ ↑↑ UC ↑ ↑ ↓↓ UC UC

ID70 ↑ UC ↓ UC UC ↓↓ UC UC

IND UC UC UC UC UC ↓ ↓ UC

BG means blood glucose concentration. UC indicates that there was no change in comparison to the values of control animals.

Out of these parameters, the hepatic FMO activity and the cytochrome b5 content of

diabetic rats showed highly significant correlations with blood glucose concentration

and consequently with each other. The correlations between FMO activity and blood

glucose level and between FMO activity and cytochrome b5 content are shown in Fig.

15. and 16.

60

350 450 5502000

3000

4000

5000

6000

7000

8000

9000

y=14.62x-2252r=0.7174n=24p<0.0001

blood glucose concentration(mg/dl)

FMO

spe

cific

act

ivity

(pm

ol*m

g-1*m

in-1

)

Fig. 15. Correlation between FMO activity and blood glucose concentration

0.1 0.2 0.3 0.4 0.5

1500

2500

3500

4500

5500

6500

7500

8500

9500

y=12490x-1081r=0.6972n=24p=0.0002

cytochrome b5 content(nmol/mg protein)

FMO

spe

cific

act

ivity

(pm

ol*m

g-1*m

in-1

)

Fig. 16. Correlation between FMO activity and cytochrome b5 content

61

The analysis also gave significant correlations between average blood glucose

concentration and hepatic CYP1A and CYP3A activities and cytochrome P450

concentration (Table 9.).

Table 9. Regression analysis between blood glucose level and cytochrome contents

and activities and flavin-containing monooxygenase activity in streptozotocin-

induced diabetic rats with or without insulin treatment

Correlated parameters n r p BG with FMO activity 24 0.7141 <0.0001 BG with total cytochrome P450 content 24 0.5655 0.004 BG with cytochrome b5 content 24 0.7381 <0.0001 BG with CYP1A activity 24 0.5957 0.0021 BG with CYP3A activity 24 0.4900 0.0151 FMO activity with cytochrome b5 content 24 0.6972 0.0002

4.2 FMO-MEDIATED METABOLISM OF CNS DRUGS

4.2.1 Tolperisone

Tolperisone undergoes not only CYP-dependent, but also CYP-independent microsomal

biotransformation to a certain extent. The use of isoform-specific cytochrome P450

inhibitors, inhibitory antibodies and experiments with recombinant P450s pointed to

CYP2D6 as the prominent enzyme in tolperisone metabolism. In addition, CYP2C19,

CYP2B6 and CYP1A2 were also involved in the metabolism to a smaller extent. The

involvement of FMO was observed using a competitive inhibitor, thiourea in HLM. The

Ki of tolperisone was determined with FMO3 recombinant supersomes using MpTS

index reaction. Since thiourea did not inhibit tolperisone metabolism and the Ki of

tolperisone was 1197 µM in the MpTS oxidase reaction, FMO was suggested to not

contribute to tolperisone metabolism. Instead, the involvement of a microsomal

reductase was assumed.

62

4.2.2 Deprenyl, methamphetamine, amphetamine

The N-oxygenated metabolites of deprenyl, methamphetamine and amphetamine

enantiomers were assessed. The drugs were incubated with recombinant human FMO1

and FMO3, and human liver microsomes, respectively. The enantioselectivity of the

substrate preference as well as the stereoselective formation of the new chiral center

upon oxidation of the prochiral tertiary nitrogen of deprenyl were observed. FMO1 was

shown to be more active in the N-oxygenation of both deprenyl and methamphetamine

isomers compared to FMO3. The deprenyl enantiomers and S-methamphetamine were

substrates of human recombinant FMO3. Conversion of amphetamine to its

hydroxylamine derivative could not be observed in incubation with either FMO1 or

FMO3. Formation of the new chiral center on the nitrogen, during N-oxidation of the

tertiary amine deprenyl, was found stereoselective. The two FMO isoforms have shown

opposite preference in the formation of this chiral center. Methamphetamine-

hydroxylamine formed from methamphetamine was further transformed by FMO.

63

5 DISCUSSION

Diabetes mellitus is a complex metabolic disorder in which the energy metabolism is

disturbed. It affects carbohydrate, lipid, protein and drug metabolism. Changes in drug

metabolizing enzymes at enzymatic and gene expression level were observed regarding

flavin-containing monooxygenase and cytochrome enzyme systems in streptozotocin

induced diabetes.

5.1 CHANGES IN THE CYTOCHROME ENZYME SYSTEM: CYTOCHROME P450 AND CYTOCHROME B5

5.1.1 Hepatic cytochrome contents and CYP isoform activities

Regarding cytochrome enzyme systems in the liver, the abundance and activity of CYPs

as well as the content of cytochrome b5 were in agreement with those described in the

literature. In diabetic state, the total cytochrome P450 and p-nitrophenol hydroxylase

activity was not affected significantly and the cytochrome b5 concentration elevated as it

was observed by Barnett113 and Ackerman110. The aminopyrine N-demethylase activity

decreased compaerably to the study of Reinke112, the ethoxyresorufin O-deethylase and

testosterone 6β-hydroxylase activities increased as described by Yamazoe114 and

Shimojo115. The formation of 6β-hydroxytestosterone is catalysed by both CYP3A and

CYP2B1.114 The lack of increase in testosterone 6β-hydroxylase activity in 70 mg/kg

streptozotocin induced (D70) rats may have been due to the marked repression of

CYP2B.

In insulin treated diabetic rats the cytochrome b5 content, ethoxyresorufin O-

deethylase and testosterone 6β-hydroxylase, but not the aminopyrine N-demethylase

activities were restored to the control value. The p-nitrophenol hydroxylase activity did

not change, but there was a major standard deviation seen which may be due to the

interindividual variation of sensitivity to insulin.

In insulin treated non-diabetic rats p-nitrophenol hydroxylase and aminopyrine

N-demethylase activities showed significant repression in comparison to control values.

Woodcroft reported that insulin suppresses CYP2E1 transcription.132 The effect of

64

insulin on CYP2E1 enzymatic activity in non-diabetic rats (IND) indicated identical

tendency.

5.2 CHANGES IN THE FMO ENZYME SYSTEM

The role of insulin and ketone bodies in the regulation of CYPs (i.e. CYP1A2,

CYP2B1, CYP2E1) is proved124

1

, ,133 134, but regarding FMO these factors have not yet

been observed.

In our study it was confirmed that the FMO activity is elevated approximately 2-

fold in streptozotocin-induced diabetic rats and FMO1 isoform is responsible for this

change. Furthermore, we have shown that FMO activity and FMO1 and FMO3 mRNA

levels were nearly restored to the control value upon insulin treatment. A considerable

decrease in blood glucose concentration was shown, which did not reach the control

value. Insulin operated only in insulin deficient state and per se did not cause any

changes in FMO status.

5.2.1 Insulin as a regulator of FMO

In the study, blood glucose level (inverse indicator of insulin concentration) was

regularly checked. Glucose is the major factor responsible for insulin secretion from

pancreas β-cells. Glucose enters these cells through GLUT-2, an insulin-independent

form of glucose transporters and induces insulin secretion indirectly.90,9

Insulin affects cell functions at gene transcriptional, posttranscriptional and

posttranslational levels. The signal transduction of insulin starts with the binding of the

ligand to its plasmamembrane receptor. The insulin receptor is only functional in a

tetrameric form, in which the monomers (2α and 2β) are connected by disulfide bonds.

Insulin binding to the extracellular α-domain of insulin receptor induces intracellular,

intramolecular autophosphorylation of tirozines on β-domain. It stimulates tirozine

kinase, which is able to carry out phosphorylation of insulin receptor substrates (IRS)

directly (without mediators). There is another main pathway of insulin mediated

signaling that includes SH2 domain (src homology domain 2) containing proteins,

which bind to the phosphorylated tirozine on the β-domain of the receptor, and become

activated. The phospholipase C (PLC-γ) enzyme possesses an SH2 domain, which

65

makes it able to bind to phosphorylated tirozines. When it becomes activated, it is

translocated to the membrane. In the lipid bilayer PIP2 (phosphatidil-inositol 4,5

biphosphate) is hydrolized by PLC-γ into inozitol 1,4,5 –triphosphate (IP3) and diacyl

glycerol (DAG). IP3 binds to the Ca-store and releases Ca2+ ions. This leads to an

intracellular Ca2+ signal. Subsequently, Ca2+ binds to calmodulin producing a Ca2+-

calmodulin complex, which in turn can bind to various enzymes and activate them.

DAG stays in the membrane and activates protein kinase C (PKC). PKC is a

serine/treonin kinase having broad substrate-specificity and it induces the raf kinase and

so the mitogen activated protein kinase (MAP kinase) cascade. The steroid receptor

coactivator (Src kinase) can also be activated in the insulin-mediated pathway. Grb2,

another protein containing SH2 domain, stays in the cytoplasm and usually binds

together with SOS, which is a helper protein in the process of transforming GDT to

GTP for ras protein. The activated ras protein stimulates raf kinase, the first enzyme in

the MAP kinase cascade activation, which leads to the phosphorylation of various

proteins, transcriptional factors and kinases.

The main pathway of insulin transduction is likely to be the protein

phosphorylation/dephosphorylation. However, insulin also acts through Ca2+, IP3 and

DAG mediators.91,135 Long-term effects of insulin mediated by gene transcription,

protein synthesis, cell proliferation and cell differentiation are known. Furthermore, as

insulin binds to the IGF-1 receptors it has a growth forming effect as well.90,91 At least

one of these signaling patways is responsible for the insulin-mediated FMO regulation.

Covalent modification could occur on IRS and PLC-γ pathways. In our study, the

changes of FMO function in diabetes and insulin treated diabetes were detected not only

at enzymatic, but also at mRNA level. Therefore, it is strongly proposed that insulin

modifies FMO gene transcription or the stability of FMO mRNA. Since the FMO

activity increased in insulin deficiency, it decreased to control level on insulin

treatment, and insulin per se did not cause any changes of the FMO status, it was

suggested that insulin possesses a repressor function. Thus, it activates those signaling

pathways that are responsible for the binding of transcriptional factors to the cis

responsive element (i.e. promoter or enchancer/silencer regions) of the FMO gene.

66

It was demonstrated that the FMO3 isoform is approximately 2 times more

sensitive to insulin concentration compared to FMO1, a fact to keep in mind since the

FMO3 isoform is the dominant one in human liver.

5.2.2 Glucose as a marker for elevated FMO activity

In our study, the insulin level was altered by insulin administration to diabetic rats

(ID70). Nevertheless, blood glucose level, the inverse parameter of insulin

concentration was regularly checked. In short-term studies the determination of glucose

concentration is simple and characteristic for the severity of diabetes mellitus, whereas

long-term studies permit the determination of an other informative parameter, namely

hemoglobin A1c (HbA1c) concentration that indicates the long-term blood glucose

concentration. Since the regression analyses showed a highly significant correlation

between FMO activity and average blood glucose concentration in diabetic rats, blood

glucose is probably a marker for elevated FMO activity. Similarly, a highly significant

correlation was observed between the FMO activity and the HbA1c level in

streptozotocin-induced diabetic rats, either treated or not with insulin and in non-

diabetic animals confirming our present result (unpublished observation). The average

blood glucose levels between the groups treated with 50 and 70 mg/kg streptozotocin

did not differ significantly implying a nearly maximal diabetogen effect of 50 mg/kg

dose. Although, the increase of FMO activity showed dose-dependence of

streptozotocin, its biological meaning is questionable.

Total cytochrome P450 content, hepatic CYP1A and CYP3A activities were also

correlated with blood glucose level in diabetic rats.

5.2.3 Ketone bodies may have a role in FMO regulation

As ketone bodies were suggested to play a role in CYP regulation, the possibility that

acetoacetate and β-hydroxy-butyrate may have an influence on FMO regulation seems

to be conceivable. It was reported that acetone concentration rose sharply along with

blood glucose concentration in rats. The concentration was below 0.4 mM until the

serum glucose level exceeded 400 mg/dl, then elevated to 6 mM between 500-

600 mg/dl glucose.11 Although the acetone concentration was not measured in this 6

67

study, our results regarding FMO activity showed that the blood glucose level of

control, ID70 and IND rats was under 400 mg/dl, accompanied by a normal FMO

activity, whereas that of D50 and D70 rats ranged between 500-600 mg/dl, accompanied

by an elevated FMO activity. These parallel changes indicate that studies on the effect

of acetone on FMO activity at such doses may provide further pieces of information.

5.2.4 FMO correlates with cytochrome b5 enzyme in experimental diabetes

Interestingly, FMO activity and cytochrome b5 content had a tendency to change in the

same manner and both parameters showed strong correlation with the blood glucose

level as well as with each other in diabetic state. The role of cytochrome b5 in fatty acid

desaturation and drug metabolism is well recognized.136 The elevation of cytochrome b5

level is accompanied by the defect of terminal desaturase enzyme in diabetes.137 As the

afore-mentioned enzymes have the tendency to change in the same direction in diabetes,

it is probable that a change of FMO activity occurs when fatty acid metabolism is

disturbed. It was shown previously that FMO and cytochrome b5 are involved in the

metabolism of amines through the catalyzes of a redox reaction in the opposite direction

(N-oxygenation and retro reduction, respectively).12 Therefore, it is reasonable to

assume a common regulatory pathway between FMO and cytochrome b5.

5.3 THE PROPOSED INFLUENCE OF DIABETES INDUCED CHANGES OF FMO ACTIVITY ON TOLPERISONE, DEPRENYL, AMPHETAMINE, METHAMPHETAMINE METABOLISM

The metabolic pathways and profiles of tolperisone, deprenyl, amphetamine and

metamphetamine were studied.

Tolperisone was found to be a poor FMO substrate in humans. Firstly, it was a

weak inhibitor of FMO3 mediated MpTS/MpTSO index reaction using recombinant

enzyme with a Ki well over the diagnostic concentration range. Secondly, the turnover

number for tolperisone biotransformation was low to be relevant. Thirdly, thiourea used

at 10 mM was not capable of inhibiting the FMO-mediated metabolism of tolperisone in

HLM. Due to the high affinity constant of FMO for tolperisone, the diabetes induced

increase in FMO capacity is suggested to not have any influence on tolperisone

metabolism.

68

Deprenyl and its metabolites (methamphetamine and amphetamine) were

metabolized by FMO efficiently to their respective N-oxides or hydroxylamines. Both

tertiary amines were primarly transformed by the FMO1 isoform, although the FMO3

mediated metabolism was also observed. Therefore, it is suggested that diabetes induced

changes of FMO activity may have influence on the metabolism of deprenyl,

methamphetamine and amphetamine.

Finally, since insulin is a regulator of FMO, I would like to emphasize the necessity to

observe diabetes induced metabolic changes regarding this enzyme system in humans. It

is vital in one hand since FMO3, the dominant FMO isoform in adult liver, appears to

be very sensitive to insulin-deficiency; and on the other hand the number of compounds

understood to become metabolized with the contribution of FMO is constantly

increasing.

69

6 CONCLUSIONS

1. Insulin has a role in the regulation of FMO. We proposed a repressor function

for insulin based on the observed insulin-deficiency induced FMO activity,

which was restored on insulin supplementation to the control level and had no

influence on non-diabetic rats. These were reported for the very first time by us.

2. We confirmed that diabetes induced the FMO enzyme approximately by 2-fold,

and observed that the insulin treatment of STZ-induced diabetic rats restored the

FMO enzymatic activity.

3. We demonstrated that diabetes had an effect on FMO function. The severity of

diabetes was characterized by blood glucose concentration. A high correlation

was shown between FMO activity and blood glucose concentration (inverse

indicator of insulin level) in diabetic rats. Blood glucose level is a good marker

for FMO induction.

4. We revealed that insulin itself did not affect either the FMO activity or the gene

expression in non-diabetic rats.

5. Furthermore, the functional changes of FMO in diabetic and insulin treated

diabetic rats were shown not only at enzymatic, but also at gene expression

level, observing FMO1 and FMO3 isoforms.

6. We have recognized that the FMO1 and the FMO3 isoforms showed distinct

sensitivity to insulin-deficiency. We have shown that the FMO3 isoform is two

times more sensitive to insulin than FMO1, under the conditions examined.

7. Our regression analysis indicated a significant correlation between FMO activity

and cytochrome b5 content.

70

7 SUMMARY

INSULIN AS A REGULATOR OF FLAVIN-CONTAINING MONOOXYGENASE

ENZYME IN STREPTOZOTOCIN-INDUCED DIABETIC RATS

The flavin-containig monooxygenase enzyme (FMO) family is one of the major

microsomal monooxygenase enzyme systems involved in drug metabolism. Its

physiological role in mammals, aside from the transformation of trimethylamine into

trimethylamine N-oxide, is unknown. FMO1 and FMO3 isoforms are predominant in

the liver of experimental animals and humans, respectively. The activity of FMO

changes in certain pathophysiological conditions, for example in diabetes. Our main

goal was to study whether insulin has a role in FMO regulation. For this purpose we

induced experimetal diabetes in rats using streptozotocin, and then the diabetic rats

received insulin supplementation. Changes in FMO function were determined at the

level of enzymatic activity and gene expression. FMO activity was measured using an

FMO specific substrate, benzydamine. The FMO1 and FMO3 gene expressions were

observed by q-RT-PCR. Along that, the changes in abundance and activities of hepatic

cytochrome enzyme system were characterized in order to support and complete the

results in our experimental model system. It was shown that both, FMO activity and

FMO1 mRNA level increased approximately 2-fold in diabetic rats. These levels were

restored to the control level upon insulin supplementation and no change was observed

upon insulin treatment of non-diabetic animals. A repressor function was proposed for

insulin, because FMO activity was induced in insulin-deficiency, restored on insulin

supplementation to control level and had no influence per se. As a high correlation was

found between the FMO activity and the blood glucose level of diabetic rats, blood

glucose level was suggested to be a good marker for elevated FMO activity.

Furthermore, we have recognized that FMO1 and FMO3 isoforms showed distinct

sensitivity to insulin-deficiency.

1. Dalmadi B, Leibinger J, Szeberényi Sz, Borbás T, Farkas S, Szombathelyi Zs and Tihanyi K (2003) Identification of metabolic pathways involved in the biotransformation of tolperisone by human microsomal enzymes. Drug Metab Dispos, 31: 631-636.

71

2. Szökő É, Tábi T, Borbás T, Dalmadi B, Tihanyi K and Magyar K. (2004) Assessment of the N-oxidation of deprenyl, methamphetamine and amphetamine enantiomers by chiral capillary electrophoresis; an in vitro metabolism study. Electrophoresis, 25(16): 2866-75.

3. Borbás T, Benkő B, Szabó I, Dalmadi B, Tihanyi K. (2006) Insulin in flavin-containing monooxygenase regulation. Flavin-containing monooxygenase and cytochrome P450 activities in experimental diabetes. Eur J Pharm Sci, 28(1-2): 51-58.

4. Borbás T, Zhang J, Cerny MA, Likó I, Cashman JR (2006) Investigation of structure and function of a catalytically efficient variant of the human flavin-containing monooxygenase form 3 (FMO3). Drug Metab and Dispos, 34: 1995-2002.

72

REFERENCES 1 Ziegler DM. Microsomal flavin-containing monooxygenase: oxygenation of

nucleophilic nitrogen and sulphur compounds. In: Jakoby WB (Ed.), Enzymatic Basis

of Detoxication. Academic Press, New York, 1980: 201-277.

2 Ziegler DM, Mitchell CH. (1972) Microsomal oxidases IV; Properties of a mixed

function amine oxidase isolated from pig liver microsomes. Arch Biochem Biophys,

150: 116-125.

3 Miller JA, Kramer JW, Miller EC. (1960) The N-and-ring hydroxylation of 2

acetylaminofluorene during carcinogenesis in the rat. Cancer Res, 20: 950-962.

4 Ziegler DM, Jollow D, Cook DF. Properties of a purified liver microsomal mixed-

function amine oxidase. In: Kamin, Henry (Ed.), Proceedings of the Third International

Symposium on Flavins and Flavin-proteins. University Park Press, Butterworth and Co

Ltd., Baltimore, London, 1971: 507-522.

5 Rettie AE, Fischer MB: Transformation Enzymes: Oxidative, Non-P450. In: Woolf TF

(Ed.), Handbook of Drug Metabolism. Marcel Dekker Inc., New York, 1999: 132-151. 6 Cashman JR. Flavin monooxygenases. In: Ioannides C. (Ed.), Enzymes systems that

Metabolise Drugs and Other Xenobiotics. John Wiley and Sons Ltd., 2002: 67-93. 7 Cashman JR. (1995) Structural and catalytic properties of the mammalian flavin-

containing monooxygenase. Chem Res Toxicol, 8(2): 165-181. 8 Brunelle A, Bi YA, Lin J, Russel B, Luy L, Berkman C, Cashman JR. (1977)

Characterization of two human flavin-containing monooxygenase (FORM3) enzymes

expressed in Escherichia coli as maltose binding protein fusions. Drug Metab Dispos,

25(8): 1001-1007.

73

9 Krueger SK, Williams DE. (2005) Mammalian flavin-containing monooxygenase:

structure/function, genetic polymorphisms and role in drug metabolism. Pharmacol

Ther, 106: 357-387. 10 Cashman JR. In vitro metabolism: FMO and related oxygenations. In: Woolf TF

(Ed.), Handbook of Drug Metabolism. Marcel Dekker Inc., Ann Arbor, 1999, 477-506. 11 Birkett DJ, MacKenzie PI, Veronese ME, Miners JO. (1993) In vitro approaches can

predict human drug metabolism. Trends Pharmacol Sci, 14(8): 292-294.

12 Cashman JR. (2005) Some distinctions between flavin-containing and cytochrome

P450 monooxygenases. Biochem Biophys Res Commun, 338: 599-604. 13 Grothusen A, Hardt J, Brautigam L, Lang D, Böcker R. (1996) A convenient method

to discriminate between cytochrome P450 enzymes and flavin-containing

monooxygenases in human liver microsomes. Arch Toxicol, 71: 64-71. 14 Poulsen LL, Hyslop RM, Ziegler DM. (1979) S-oxygenation of N-substituted

thioureas catalysed by the pig liver microsomal FAD-containing monooxygenase. Arch

Biochem Biophys, 198: 78-88.

15 Larsen-Su S, Williams DE. (1996) Dietary indole-3-carbinol inhibits FMO activity

and the expression of flavin-containing monooxygenase form 1 in rat liver and intestine.

Drug Metab Dispos, 24(9): 927-931.

16 Correira MA. Human and rat liver cytochromes P450: Functional markers, diagnostic

inhibitor probes, and parameters frequently used in P450 studies (appendix). In: Ortiz

de Montellano PR (Ed.), Cytochrome P450 structure, mechanism and biochemistry.

Kluwer Academic/Plenum Publishers, New York, 2005: 619-943. 17 Bhagwat SV, Bhamre S, Boyd MR, Ravindranath V. (1996) Further Characterization

of rat brain flavin-containing monooxygenase. Metabolism of imipramine to its N-

oxide. Biochem Pharmacol, 51(11): 1469-75.

74

18 Myers CR, Porgilsson B, Myers JM.(1997) Antibobies to a synthetic peptide that

react with flavin-containing monooxygenase (HLFMO3) in human hepatic microsomes.

J Pharmacol Toxicol Methods, 37(2): 61-66.

19 Pike MG, Mays DC, Macomber DW, Lipsky JJ. (2001) Metabolism of disulfiram

metabolite, S-methyl N,N-diethyldithiocarbamate, by flavin monooxygenase in human

renal microsomes. Drug Metab Dispos, 29: 127-132.

20 Engler H, Taurog A, Nakashima T. (1982) Mechanism of inactivation of thyroid

peroxidase by thioureylene drugs. Biochem Pharmacol, 31: 3801-3806.

21 Kedderis GL, Rickert DE. (1985) Loss of rat liver microsomal cytochrome P450

during methimazole metabolism. Role of flavin-containing monooxygenase. Drug

Metab Dispos, 13: 58-61.

22 Katchamart S, Stresser DM, Dehal SS, Kupfer D, Williams DE. (2000) Concurrent

flavin-containing monooxygenase down-regulation and cytochrome P-450 induction by

dietary indoles in rat: implications for drug-drug interactions. Drug Metab Dispos, 28:

930-936. 23 Lawton M, Cashman J, Cresteil T, Dolphin C, Alfarra A. (1994) A nomenclature for

the mammalian flavin-containing monooxygenase gene family based on amino acid

sequence identities. Arch Biochem Biophys, 308: 254-257.

24 Hines RN, Hopp KA, Franco J, Saeian K, Begun FP. (2002) Alternative processing of

the human FMO6 gene renders transcripts incapable of encoding a functional flavin-

containing monooxygenase. Mol Pharmacol, 62: 320-25.

25 Philips IR, Dolphin CT, Clair P, Hadley MR, Hutt AJ. (1995) The molecular biology

of the flavin-containing monooxygenase of man. Chem Biol Interact, 96(1): 17-32.

75

26 Shephard EA, Dolphin CT, Fox ME, Povey S, Smith R, Phillips IR. (1993)

Localization of genes encoding three distinct flavin-containing monooxygenase to

human chromosome 1g. Genomics, 16: 85-89.

27 Dolphin CT, Riley JH, Smith RL, Shephard EA, Phillips IR. (1997) Structural

organization of the human flavin-containing monooxygenase 3 gene (FMO3), the

flavored candidate for Fish Odor Syndrome, determined directly from genomic DNA.

Genomics, 46: 260-267.

28 McCombie RR, Dolphin CT, Povey S, Phillips IR, Shepard E. (1996) Localization of

human flavin-containing monooxygenase genes FMO2 and FMO5 to chromosome 1q.

Genomics, 1996, 34(3):426-9.

29 Wu RF, Ichikawa Y. (1994) Characteristic properties and kinetic amnalysis with

neurotoxins of porcine FAD-containing monooxygenase. Biochem Biophys Acta, 1208:

204-210.

30 Ziegler DM. Flavin-containing monooxygenase family of isozymes. Abstract, Proc of

the First International Workshop on Trimethylaminuria, Bethesda, MD, March 29-30.,

1999.

31 Ziegler DM. (2002) An overview of mechanism, substrate specificities and structure

of FMOs. Drug Metab Rev, 34(3): 503-511. 32 Yeung CK, Lang DH, Thummel KE, Rettie AE. (2000) Immunoquantitation of

FMO1 in human liver, kidney and intestine. Drug Metab Dispos, 28: 1107-1111.

33 Overby LH, Carver GC, Philpot RM. (1997) Quantitation and kinetic properties of

hepatic microsomal and recombinant flavin-containing monooxygenases 3 and 5 from

humans. Chem Biol Interact, 106: 29-45.

76

34 Cashman JR, Camp K, Fakharzadeh SS, Fenessey PV, Hines RN, Mamer OA,

Mitchell SC, Preti G, Schlenk D, Smith RL, Tjoa SS, Williams DE, Yannicelli S. (2003)

Biochemical and clinical aspects of the human flavin-containing monooxygenase form

3 (FMO3) related to trimethylaminuria. Curr Drug Metab, 4: 151-170.

35 Zhang J, Cashman JR. (2006) Quantitative analysis of FMO gene mRNA levels in

human tissues. Drug Metab Dispos, 34: 19-26.

36 Cashman JR. (2000) Human flavin-containing monooxygenase: substrate specificity

and role in drug metabolism. Curr Drug Metab, 1(2): 181-191. 37 Ripp SL, Itagaki K, Philpot RM, Elfarra AA. (1999) Species and sex differences in

expression of flavin-containing monooxygenase form 3 in liver and kidney microsomes.

Drug Metab Dispos, 27(1): 46-52.

38 Cherrington NJ, Cao Y, Cherrington JW, Rose RL, Hodgson E. (1998) Physiological

factors affecting protein expression of flavin-containing monooxygenases 1, 3 and 5.

Xenobiotica, 28(7): 673-682.

39 Kim YM, Ziegler DM. (2000) Size limit of thiocarbamides accepted as substrates by

human flavin-containing monooxygenase 1. Drug Metab Dispos, 28: 1003-1006.

40 Cashman JR and Zhang J. (2006) Human flavin-monooxygenases. Annu Rev

Pharmacol Toxicol, 46: 65-100. 41 Lin J, Cashman JR. (1997) Detoxication of tyramine by the flavin-monooxygenase:

stereoselective formation of the trans oxime. Chem Res Toxicol, 10: 842-52.

42 Yamada H, Baba T, Oguri K, Yoshimura H. (1988) The mammalian flavin-

containing monooxygenases: molecular characterization and regulation of expression.

Biochem Pharmacol, 37(2): 368-370.

77

43 Lin J, Cashman JR. (1997) N-oxygenation of phenethylamine to the trans-oxime by

adult human liver flavin-containing monooxygenase and retroreduction of

phenethylamine hydroxylamine by human liver microsomes. J Pharmacol Exp Ther,

282(3): 1269-79.

44 Hadley MR, Svajdlenka E, Damani LA, Oldham HG, Tribe J, Camilleri P, Hutt AJ.

(1994) Species variability in the stereoselective N-oxidation of pargyline. Chilarity,

6(2): 91-7.

45 Kashiyama E, Yokoi T, Odomi M, Kamataki T. (1999) Stereoselective S-oxidation

and reduction of flosequinan in rat. Xenobiotica, 29(8): 815-26.

46 Ripp SL, Itagaki K, Philpot RM, Elfarra AA. (1999) Methionine S-oxidation in

human and rabbit liver microsomes: evidence for a high-affinity methionine S-oxidase

activity that is distinct from flavin-containing monooxygenase 3. Arch Biochem

Biophys, 367: 322-332.

47 Hines RN, Cashman JR, Philpot RM, Williams DE, Ziegler DM. (1994) The

mammalian flavin-containing monooxygenases: molecular characterization and

regulation of expression. Toxicol Appl Pharmacol, 125(1): 1-6.

48 Cashman JR, Akerman BR, Forrest SM, Treacy EP. (2000) Population-specific

polymorphisms of the human FMO3 gene: significance for detoxication. Drug Metab

Dispos, 28(2): 169-173.

49 Lang DH, Rettie AE. (2000) In vitro evaluation of potential in vivo probes for human

flavin-containing monooxygenase (FMO): metabolism of benzydamine and caffeine by

FMO and P450 isoforms. Br J Clin Pharmacol, 50: 311-314. 50 Kawaji A, Ohara K, Takabatake E. (1993) An assay of flavin-containing

monooxygenase activity with benzydamine N-oxidation. Anal Biochem, 214: 409-412.

78

51 Störmer E, Roots I, Brockmöller J. (2000) Benzydamine N-oxydation as an index

reaction reflecting FMO activity in human liver microsomes and impact of FMO3

polymorphisms on enzyme activity. Br J Clin Pharmacol, 50: 553-561.

52 Chasseaud LF, Catanese B. (1985) Pharmacokinetics of benzydamine. Int J Tissue

React, 7: 195-204. 53 Park CS, Chung WG, Kang JH, Roh HK, Lee KH, Cha YN. (1999) Phenotyping of

flavin-containing monooxygenase using caffeine metabolism and genotyping of FMO3

gene in a Korean population. Pharmacogenetics, 9(2): 155-164.

54 Bjeldanes LF, Kim JY, Grose KR, Bartholomew JC, Bradfield CA. (1991) Aromatic

hydrocarbon responsivness-receptor agonists generated from indole-3-carbinol in in

vitro and in in vivo:comparison with 2,3,7,8-tetrachloro-dibenzo-p-dioxin. Proc Natl

Acad Sci USA, 88: 9543-9547.

55 Cashman JR, Xiong Y, Lin J, Verhagen H, Poppel G, Bladeren PJ, Larsen-Su S,

Willimams DE. (1999) In vitro and in vivo inhibition of human flavin-containing

monooxygenase form 3 (FMO3) in the presence of dietary indoles. Biochem Pharmacol,

58: 1047-1055.

56 Katchamart S, Williams DE. (2001) Indole-3-modulation of hepatic monooxygenases

CYP1A1, CYP1A2 and FMO1 in guinea pig, mouse and rabbit. Comp Biochem Physiol

C Toxicol Pharmacol, 129: 377-384.

57 Ziegler DM. (1988) Flavin-containing monooxygenases: catalytic mechanism and

substrate specificities. Drug Metab Rev, 19: 1-32.

58 Ziegler DM. (1990) Flavin-containing monooxygenases: enzymes adapted for

multisubstrate specificity. Trends Pharmacol Sci, 11(8): 321-324.

79

59 Zhao Y, Christensen, Fankhauser C, Cashman JR, Cohen JD, Weigel D, Chory J.

(2001) A role for flavin monooxygenase-like enzymes in auxin biosynthesis. Science,

291: 306-309.

60 Suh JK, Poulsen LL, Ziegler DM, Robertus JD. (1999) Yeast flavin-containing

monooxygenase generates oxidizing equivalents controlling protein folding in the

endoplasmatic reticulum. Proc Natl Acad Sci USA, 96: 2687-2691.

61 Suh J, Robertus JD. (2000) Yeast FMO is induced by the unfolded protein response.

Proc. Natl. Acad. Sci. USA, 2000, 97:121-126.

62 Ziegler DM, Duffel MW, Poulsen LL. (1979) Studies on the nature and regulation of

the cellular thio:disulphide potential. Ciba Found Symp, 72: 191-204.

63 Jeitner TM, Lawrence DA. (2001) Mechanisms for the cytotoxicity of cysteamine.

Toxicol Sci, 62: 57-64.

64 Khomenko T, Deng X, Sandor Z, Tarnawski AS, Szabo S. (2004) Cystemanine alters

redox state, HIF-1α transcriptional interactions and reduces duodenal mucosal

oxygenation: novel insight into the mechanisms of duodenal ulceration. Biochem

Biophys Res Commun, 317: 121-127.

65 Hagen TM, Moreau R, Suh JH, Visioli F. (2002) Mitochondrial decay in the aging rat

heart: evidence for improvement by dietary supplementation with acetyl-carnitine

and/or lipoic acid. Ann N Y Acad Sci, 959: 491-507.

66 Cerny MA, Hanzlik RP. (2005) Cyclopropylamine in action of cytochrome P450: role

of metabolite intermediate complexes. Biochem Biophys, 436: 265-275.

67 Rodriguez RJ, Buckholz CS. (2003) Hepatotoxicity of ketoconazole in Sprague

Dawley rats: glutathione depletion, flavin-containing monooxygenases-mediated

bioactivation and hepatic covalent binding. Xenobiotica, 33: 429-441.

80

68 Cashman JR. (2003) The role of flavin-containing monooxygenases in drug

metabolism and development. Curr Opin Drug Discov Devel, 6(4): 486-493.

69 Ayesh R, Smith RL. (1990) Genetic polymorphism of trimethylamine N-oxidation.

Pharmacol Ther, 45: 387-401.

70 Treacy EP, Akerman BR, Chow LM, Youil R, Bibeau C, Lin J. (1998) Mutations of

the flavin-containing monooxygenase gene (FMO3) cause trimethylaminuria, a defect in

detoxication. Hum Mol Genet, 7: 839-45.

71 Mitchell SC, Smith RL. (2001) Trimethylaminuria: the fish malodor syndrome. Drug

Metab Dispos, 29: 517-521.

72 Humbert JR, Hammond KB, Hathaway WE, Marcoux JG, O’Brien D. (1970)

Trimethylaminuria: fish-odour syndrome. The Lancet, 4: 970-971. 73 Mitchell SC, Smith RL. (2003) Trimethylamine and odorous sweat. J Inherit Metab

Dis, 26: 415-16.

74 Mitchell SC, Zhang AQ, Barrett T, Ayesh R, Smith RL. (1997) Studies on the

dicontinous N-oxidation of trimethylamine among Jordanian, Ecuadorian and New-

Guinean populations. Pharmacogenetics, 7: 45-50.

75 McConnell HW, Mitchel SC, Smith RL, Brewster M. (1997) Trimethylaminuria

associated with seizures and behavioural disturbance: a case riport. Seizure, 6: 317-322.

76 Luo Z, Hines RN. (2001) Regulation of flavin-containing monooxygenase 1

expression by Ying Yang and Hepatic Nuclear Factors 1 and 4. Mol Pharmacol, 60:

1421-1430.

81

77 Luo Z, Hines RN. (1997) Further characterization of the major and minor rabbit

FMO1 promoters and identification of both positive and negative distal regulatory

elements. Arch Biochem Biophys, 346: 96-104.

78 El-Alfy, Larsen B, Schlenk D. (2002) Effect of cortisol and urea on flavin

monooxygenase activity and expression in rainbow trout, Oncorhynchus mykiss. Mar

Environ Res, 54: 275-278. 79 Ryu SD, Yi HG, Cha YN, Kang JH, Kang JS, Jeon YC, Park HK, Yu TM, Lee JN,

Park CS. (2004) Flavin-containing monooxygenase activity can be inhibited by nitric

oxide-mediated S-nitrosylation. Life Sci, 75(21): 2559-72. 80 Kaderlik RK, Weser E, Ziegler DM. (1991) Selective loss of liver flavin-containing

monooxygenase in rats on chemically defined diets. Prog Pharm Clin Pharm, 8(3): 95-

103.

81 Cashman JR, Lattard V, Lin J. (2004) Effect of total parenteral nutrition and choline

on hepatic flavin-containing and cytochrome P450 monooxygenase activity in rats.

Drug Metab Dispos, 32: 222-229. 82 Dixit A and Roche TE. (1984) Spectrophotometric assay of the flavin-containing

monooxygenase and changes in its activity in female mouse liver with nutritional and

diurnal conditions. Archs Biochem Biophys, 238 (1): 50-63.

83 Brodfuehrer JI and Zannoni VG. (1986) Ascorbic acid deficiency and flavin-

containing monooxygenase. Bichem Pharmacol, 35(4): 637-644.

84 Coecke S, Debast G, Phillips IR, Vercruysse A, Shepard EA, Rogiers V. (1998)

Hormonal regulation of microsomal flavin-containing monooxygenase activity by sex

steroids and growth hormone in co-cultured adult male rat hepatocytes. Biochem

Pharmacol, 56: 1047-1051.

82

85 Dannan GA, Guengerich FP, Waxman DJ. (1986) Hormonal regulation of rat liver

microsomal enzymes. J Biol Chem, 261: 10728-10735. 86 Hines RN, Cashman JR, Philpot RM, Williams DE, Ziegler DM. (1994) The

mammalian flavin-containing monooxygenases: molecular characterisation and

regulation of expression. Toxicol Appl Pharmacol, 125: 1-6.

87 Osimitz TG, Kulkarni AP. (1982) Oxidative metabolism of xenobiotics during

pregnancy. Significance of microsomal flavin-containing monooxygenase. Biochem

Biophys Res Commun, 109: 1164-1171.

88 Heinze E, Hlavica P and Kiese M. (1970) N-oxygenation of arylamines in

microsomes prepared from corpra lutea of the cycle and other tissues of the pig.

Biochem Pharmacol, 19: 641-649.

89 Roderick E. McGrew. Diabetes Mellitus. In: Roderick E. McGrew (Ed.),

Encyclopedia of medical history. Macmillan Press, London, 1985: 90-93.

90 Ádám V, Mandl J. A szénhidrátok anyagcseréje. In: Ádám V (Ed.), Orvosi biokémia.

Semmelweis Kiadó, Budapest, 1996: 76-120.

91 Tímár J. A szénhidrát-anyagcsere gyógyszertana. In: Fürst Zs (Ed.), Gyógyszertan.

Medicina Könyvkiadó Rt., Budapest, 1999: 742-757.

92 Dainith H, Stevenson IH, O’Malley K. (1976) Influence of diabetes mellitus on drug

metabolism in man. Int J Clin Pharmacol, 13: 55-58.

93 Adithan C, Danda D, Swaminathan RP, Indhiresan J, Shasindran CH, Bapna JS,

Chandrasekar S. (1988) Effect of type I diabetes mellitus on theophylline elimination.

Med Sci Res, 16: 427-428.

83

94 Adithan C, Sriram G, Swaminathan RP, Krishnan M, Bapna JS, Chandrasekar S.

(1989) Effect of type II diabetes mellitus on theophylline elimination. Int J Clin

Pharmacol Ther Toxicol, 27: 258-260.

95 Gwilt PR, Nahhas RR, Tracewell WG. (1991) The effects of diabetes mellitus on

pharmacokinetics and pharmacodynamics in humans. Clin Pharmacokinet, 20(6): 477-

490.

96 Adithan C, Danda D, Shasindran CH, Bapna JS, Swaminathan RP, Chandrasekar S.

(1989) Differential effect of type I and type II diabetes mellitus on antyipyrine

elimination. Meth and Find Exp Clin Pharmacol, 11(12): 755-758.

97 Matzke GR, Frye RF, Early JJ, Straka RJ, Carson SW. (2000) Evaluation of the

influence of diabetes mellitus on antipyrine metabolism and CYP1A2 and CYP2D6

activity. Pharmacotherapy, 20(2): 182-190.

98 Dajani RM, Kayyali S, Saheb SE, Birbari A. (1974) A study on the physiological

disposition of acetophenetidin by the diabetic man. Comp Gen Pharmacol, 5:1-9.

99 Sotaniemi EA, Pelkonen O, Arranto AJ, Tapanainen P, Rautio A, Pasanen M. (2002)

Diabetes and elimination of antipyrine in man: an analysis of 298 patients classified by

type of diabetes, age, sex, duration of disease and liver involvement. Pharmacol

Toxicol, 90: 155-160.

100 Zysset T, Wiwtholtz H. (1988) Differential effect of type I and type II diabetes on

antipyrine disposition in man. Eur J Clin Pharmacol, 34: 369-375.

101 Takamura T, Shakurai M, Ota T, Ando H, Honda M, Kaneko S. (2004) Genes for

systemic vascular complications are differentially expressed in the livers of type 2

diabetic patients. Diabetologia, 47: 638-647.

84

102 Elsner M, Guldbakke B, Tiedge M, Munday R, Lenzen S. (2000) Relative

importance of transport and alkylation for pancreatic β-cell toxicity of streptozotocin.

Diabetologia, 43: 1528-1533.

103 Gruppuso PA, Boylan JM, Posner BI, Faure R, Braitigan DL. (1990) Hepatic protein

phosphotyrosine phosphatase. Dephosphorylation of insulin receptor and epidermal

growth factor receptors in normal and alloxan diabetic rats. J Clin Invest, 85: 1754-

1760.

104 Dunn JS, Sheehan HL, McLethie NGB. (1943) Necrosis of islets of Langerhans

produced experimentally. Lancet, 1: 484-487.

105 Arison RN, Ciaccio EI, Glitzer MS, Cassaro JA, Pruss MP. (1967) Light and

electron microscopy of lesions in rats rendered diabetic with streptozotocin. Diabetes,

16: 51-56.

106 Dixon RL, Hart LG, Foust JR. (1961) The metabolism of drugs by liver microsomes

from alloxan diabetic rats. J Pharmacol Exp Ther, 133: 7-11.

107 Dajani RM and Saheb SE. (1971) In vitro and in vivo studies of the metabolism of

phenylbutazone in the alloxan rat and rabbit (Abstract). J Pharm Pharmacol, (Suppl) 23:

220S-221S.

108 Nishihata T, Yata N, Kamada A. (1979) Role of acetone bodies in the abnormal

pharmacokinetic behaviour of chlorpropramide in alloxan diabetic rabbits. Chem Pharm

Bull, 27: 1740-1746.

109 Dixon RL, Hart LG, Rogers LA, Fouts JR. (1963) The metabolism of drugs by liver

microsomes from alloxan diabetic rats: long-term diabetes. J Pharmacol Exp Ther, 142:

312- 317.

85

110 Ackerman DM and Leibman KC. (1977) Effect of experimental diabates on drug

metabolism in the rat. Drug Metab Dispos, 5(4): 405-410.

111 Rouer E, Lemoine A, Cresteil P, Rouet P, Leroux JP. (1987) Effects of genetic or

chemically induced diabetes on imipramine metabolism. Drug Metab Dispos, 15(4):

524-528.

112 Reinke LA, Stohs SJ, Rosenberg H. (1978) Altered activity of hepatic mixed-

function monooxygenase enzymes in streptozotocin-induced diabetic rats. Xenobiotica,

8(10): 611-619.

113 Barnett CR, Flatt PR, Ioannides C. (1994) Modulation of the rat hepatic cytochrome

P450 composition by long-term streptozotocin-induced insulin-dependent diabetes. J

Biochem Toxicol, 9(2): 63-69.

114 Yamazoe Y, Murayama N, Shimada M, Yamauchi K, Kato R. (1989) Cytochrome

P450 in livers of diabetic rats: regulation by growth hormone and insulin. Arch

Biochem Biophys, 268: 567-575.

115 Shimojo N, Ishizaki T, Imaoka S, Funae Y, Fujii S, Okuda K. (1993) Changes in

amounts of cytochrome P450 isozymes and levels of catalytic activities in hepatic and

renal microsomes of rats with streptozotocin-induced diabetes. Biochem Pharmacol,

46(4): 621-627.

116 Schenkman JB. (1991) Induction of diabetes and evaluation of diabetic state on P450

expression. Methods Enzymol, 206: 325-331.

117 Rouer E and Leroux JP. (1980) Liver microsomal cytochrome P-450 related

monooxygenase activities is genetically hyperglycemic (ob/ob and db/db) and lean

streptozotocin-treated mice. Biochem Pharmacol, 29: 1959-1962.

86

118 Rouer E, Rouet P, Delpech M, Leroux JP. (1988) Purification and comparison of

liver microsomal flavin-containing monooxygenase from normal and streptozotocin-

diabetic rats. Biocheml Pharmacol, 37(18): 3455-3459.

119 Wang T, Shankar K, Ronis MJJ, Mehendale HM. (2000) Potentiation of

thioacetamide liver injury in diabetic rats is due to induced CYP2E1. J Pharmacol Exp

Ther 294: 473-479.

120 Cavagnaro J, Rauckman EJ, Rosen GM. (1981) Estimation of FAD monooxygenase

in microsomal preparations. Anal Biochem, 118: 204-211.

121 Holloway CT, Garfield SA. (1981) Effect of diabetes and insulin replacement on the

lipid on properties of hepatic smooth endoplasmic reticulum. Lipids, 16: 525-532.

122 Rouer E, Lemoine A, Cresteil P, Rouet P, Leroux JP. (1987) Effects of genetic or

chemically induced diabetes on imipramine metabolism. Drug Metab Dispos, 15(4):

524-528.

123 Barnett CR, Wilson J, Wolf CR, Flatt PR, Ioannides C. (1992) Hyperinsulinaemia

causes a preferential increase in hepatic P4501A2 activity. Biochem Pharmacol, 43(6):

1255-1261.

124 Woodcroft KJ, Novak RF. (1999) Insulin differentially affects xenobiotic-

enchanced, cytochrome P450 (CYP)2E1, CYP2B, CYP3A and CYP4A expression in

primary cultured rat hepatocytes. J Pharmacol Exp Ther, 289: 1121-1127.

125 Pike MG, Martin YN, Mays DC, Benson LM, Naylor S, Lipsky JJ. (1999) Roles of

FMO and P450 in the metabolism in human liver microsomes of S-methyl-N,N-

diethyldithiocarbamate, a dilsulfiram metabolite. Alcohol Clin Exp Res, 23(7): 1173-

1179.

87

126 Lowry OH, Rosebrough NJ, Farr AL and Randall RJ. (1951) Protein measurement

with the folin phenol reagent. J Biol Chem, 193(1): 265-275.

127 Greim H. (1970) Synthesesteigerung und abbauhemmung bei der vermehrung der

mikrosomalen cytochrome P450 und b-5 durch pehenobarbital. Naunyn-Schmiedebergs

Arch Pharmak, 266: 261-275.

128 Burke MD. (1985) Ethoxy-, pentoxy- and benzyloxyphenoxazones and homologues:

A series of substrates distinguish between different induced cytochromes P450.

Biochem Pharmacol, 34: 33-37.

129 Nash T. (1953) The colorimetric estimation of formaldehyde by means of the

Hantzsch reaction. Biochem J, 55(3): 416-21.

130 Reinke LA, Moyer MJ. (1985) Para-nitrophenol hydroxylation: a microsomal

oxidation which is highly inducible by ethanol. Drug Metab Dispos, 13: 548-552.

131 Anderson CD, Wang J, Kumar JN, McMilan JM, Walle UK, Walle T. (1995)

Dexamethasone induction of taxol metabolism in the rat. Drug Metab Dispos, 23: 1286-

1291.

132 Woodcroft KJ, Hafner MS, Novak RF. (2002) Insulin signaling in the transcriptional

and posttranscriptional regulation of CYP2E1 expression. Hepatology, 35(2): 263-273.

133 Barnett CR, Gibson GG, Wolf CR, Flatt PR, Ioannides C. (1990) Induction of

cytochrome P450III and P450IV family proteins in streptozotocin-induced diabetes.

Biochem J, 286: 765-769.

134 Bellward GD, Chang T, Rodriguez B, McNeill JH, Mains S, Ryan DE, Levin W,

Thomas PE. (1988) Hepatic cytochrome P450 induction in the spontaneously diabetic

BB rat. Mol Pharmacol, 33: 140-143.

88

135 Ádám V, Faragó A. Extracelluláris jelek receptorai és a jelátviteli mechanizmusok.

In: Ádám V (Ed.), Orvosi biokémia. Semmelweis Kiadó, Budapest, 1996: 371-415.

136 Schenkman JB, Jansson I. (2003) The many roles of cytochrome b5. Pharmacol Ther,

97(2): 139-152.

137 Eck MG, Wynn JO, Carter WJ, Faas FH. (1979) Fatty acid desaturation in

experimental diabetes mellitus. Diabetes, 28(5): 479-485.

89

8 PUBLICATIONS Publications the dissertation based on

1. Balázs Dalmadi, János Leibinger, Szabolcs Szeberényi, Tímea Borbás, Sándor

Farkas, Zsolt Szombathelyi and Károly Tihanyi: Identification of metabolic

pathways involved in the biotransformation of tolperisone by human microsomal

enzymes. Drug Metabolism and Disposition, 31:631-636, 2003

2. Éva Szökő, Tamás Tábi, Tímea Borbás, Balázs Dalmadi, Károly Tihanyi and

Kálmán Magyar: Assessment of the N-oxidation of deprenyl, methamphetamine and

amphetamine enantiomers by chiral capillary electrophoresis; an in vitro metabolism

study. Electrophoresis, 25(16): 2866-75, 2004

3. Tímea Borbás, Bernadett Benkő, Imola Szabó, Balázs Dalmadi, Károly Tihanyi:

Insulin in flavin-monooxygenase regulation. Flavin-containing monooxygenase and

cytochrome P450 activities in experimental diabetes. European Journal of

Pharmaceutical Sciences, 28(1-2): 51-58, 2006

4. Borbás T, Zhang J, Cerny MA, Likó I, Cashman JR: Investigation of structure and

function of a catalytically efficient variant of the human flavin-containing

monooxygenase form 3 (FMO3). Drug Metabolism and Disposition, 34: 1995-2002,

2006

Other publications

1. Borbás Tímea, Hegyesi Hargita: Huntington chorea: genetika és biokémia,

diagnosztika és terápia. Acta Pharmaceutica Hungarica, 72(2):127-36, 2002

Oral presentations

1. Dalmadi Balázs, Leibinger János, Szeberényi Szabolcs, Borbás Tímea, Pásztor

Gabriella, Farkas Sándor, Tihanyi Károly: A tolperisone in-vitro metabolikus

vizsgálata. 30 ÉVES JUBILEUMI FARMAKOKINETIKA ÉS

GYÓGYSZERMETABOLIZMUS SZIMPÓZIUM, 2002. április 4-6., Mátraháza

2. Dalmadi Balázs, Leibinger János, Szeberényi Szabolcs, Borbás Tímea, Pásztor

Gabriella, Farkas Sándor, Tihanyi Károly: A tolperisone in-vitro metabolikus

vizsgálata. VI. CLAUDER OTTÓ EMLÉKVERSENY, 2002. szeptember 27-28.,

Budapest

90

3. Borbás Tímea, Dalmadi Balázs, Szeberényi Szabolcs, Leibinger János, Beke Gyula,

Tihanyi Károly: Korreláció az egyes citokróm P450 izoenzimek és a flavin-

monooxigenáz aktivitása között. VI. CLAUDER OTTÓ EMLÉKVERSENY, 2002.

szeptember 27-28., Budapest

4. Borbás Tímea, Benkő Bernadett, Dalmadi Balázs, Szeberényi Szabolcs, Leibinger

János, Beke Gyula, Tihanyi Károly: Koexpresszió CYP és FMO izoformák között

enzimatikus szinten, humán és patkány mikroszómán vizsgálva. Ph.D.

TUDOMÁNYOS NAPOK 2003, 2003. április 10-11., Budapest

5. Benkő Bernadett, Borbás Tímea, Dalmadi Balázs, Szeberényi Szabolcs, Leibinger

János, Tihanyi Károly: Intesztinális gyógyszermetabolizmus jelentősége és stabil

citokróm P450 tartalmú mikroszóma preparálása patkány vékonybélből. Ph.D.

TUDOMÁNYOS NAPOK 2003, 2003. április 10-11., Budapest

6. Rapavi Erika, Szeberényi Szabolcs, Borbás Tímea, Lugasi Andrea, Szentmihályi

Klára, Pallai Zsolt, Kurucz Tímea, Kocsis Ibolya, Taba Gabriella, Blázovics Anna:

Természetes eredetű gyógyszer hatása a szöveti redox-státuszra patkányban.

MAGYAR SZABADGYÖK-KUTATÓ TÁRSASÁG II. KONFERENCIÁJA, 2003.

szeptember 25-27., Sopron

7. Rapavi Erika, Kocsis Ibolya, Szentmihályi Klára, Lugasi Andrea, Fehér Erzsébet,

Bányai Éva, Rhenzo Gonzalez-Cabello, Székely Edit, Szeberényi Szabolcs, Borbás

Tímea, Pallai Zsolt, Kurucz Tímea, Balázs Andrea, Czinner Erika, Héthelyi Éva,

Hagymási Krisztina, Bárkovics Sarolta, Pintér Edina, Bíró Erzsébet, Fehér János,

Blázovics Anna: Újabb adatok a természetes hatóanyagok in vitro és in vivo

hatásmechanizmusának megértéséhez. SZÉCHENYI SZIMPÓZIUM, 2004. március

10., Budapest

8. Borbás Tímea, Benkő Bernadett és Tihanyi Károly: Streptozotocinnal kiváltott

diabétesz hatása a hepatikus gyógyszermetabolizmusra patkányban. Ph.D.

TUDOMÁNYOS NAPOK 2004. április 8-9., Budapest

9. Benkő Bernadett, Borbás Tímea és Tihanyi Károly: Streptozotocinnal kiváltott

diabétesz hatása az intesztinális gyógyszermetabolizmusra patkányban. Ph.D.

TUDOMÁNYOS NAPOK 2004. április 8-9., Budapest

91

10. Benkő Bernadett, Borbás Tímea, Győrke Imola: Streptozotocin-indukálta diabétesz

hatása az intesztinális és hepatikus gyógyszermetabolizmusra patkányban. VII.

CLAUDER OTTÓ EMLÉKVERSENY, 2004. október 14-15., Budapest

11. Borbás Tímea: Flavin-monooxigenáz (HBRI). FARMAKOKINETIKA KLUB,

2005. március 2., Budapest

Posters

1. Tímea Borbás, Balázs Dalmadi, János Leibinger, Szabolcs Szeberényi, Zsolt

Szombathelyi and Károly Tihanyi: Co-expression of CYP and FMO isoforms based

on enzymatic activity in rat and in human liver microsomes. 8. EUROPEAN

INTERNATIONAL SOCIETY FOR THE STUDY OF XENOBIOTICS MEETING, 27.

April - 01. May 2003.

2. Balázs Dalmadi, János Leibinger, Szabolcs Szeberényi, Tímea Borbás, Zsolt

Szombathelyi and Károly Tihanyi: Kinetic characterization of tolperisone

hydroxylation by human microsomal enzymes. 8. EUROPEAN INTERNATIONAL

SOCIETY FOR THE STUDY OF XENOBIOTICS MEETING-ISSX, 27. April - 01.

May 2003.

3. Borbás Tímea, Benkő Bernadett, Dalmadi Balázs, Győrke Imola, Vastag Mónika és

Tihanyi Károly: Humán hepatocita citokróm P450 izoenzimek adatainak statisztikai

kiértékelése. GYÓGYSZER AZ EZREDFORDULÓN V. TOVÁBBKÉPZŐ

KONFERENCIA, 2004 március 25-27., Sopron

4. Borbás Tímea, Benkő Bernadett, Galgóczy Kornél, Dalmadi Balázs, Győrke Imola

és Tihanyi Károly: Streptozotocin-indukálta diabétesz hatása a hepatikus és

intesztinális gyógyszermetabolizmusra patkányban. FARMAKOKINETIKAI ÉS

GYÓGYSZERMETABOLIZMUS TOVÁBBKÉPZŐ SZIMPÓZIUM, 2004. április

15-17., Mátraháza

5. Borbás Tímea, Benkő Bernadett, Dalmadi Balázs, Győrke Imola, Vastag Mónika és

Tihanyi Károly: Humán hepatocita citokróm P450 izoenzimek adatainak statisztikai

kiértékelése. FARMAKOKINETIKAI ÉS GYÓGYSZERMETABOLIZMUS

TOVÁBBKÉPZŐ SZIMPÓZIUM, 2004. április 15-17., Mátraháza

6. Rapavi Erika, Szeberényi Szabolcs, Borbás Tímea, Kocsis Ibolya, Fehér Erzsébet,

Lugasi Andrea, Blázovics Anna: Citrus flavonoidok hatása a drogmetabolizáló

92

enzimrendszerre és a szöveti antioxidáns kapacitásra eltérő eredetű

májkárosodásban, patkányban. Ph.D. TUDOMÁNYOS NAPOK 2004. április 8-9.,

Budapest

7. Borbás Tímea, Jun Zhang, Matt A. Cerny, István Likó, John Cashman: Investigation

of structure and function of a catalitically efficient variant of the human flavin-

containing monooxygenase form 3 (FMO3). FARMAKOKINETIKAI ÉS

GYÓGYSZERMETABOLIZMUS TOVÁBBKÉPZŐ SZIMPÓZIUM, 2006. április

19-21., Mátraháza

8. Borbás Tímea, Jun Zhang, Matt A. Cerny, István Likó, John Cashman: Investigation

of structure and function of a catalitically efficient variant of the human flavin-

containing monooxygenase form 3 (FMO3). A MAGYAR FARMAKOLÓGIAI

TÁRSASÁG EXPERIMENTÁLIS SZEKCIÓJÁNAK II. VÁNDORGYŰLÉSE,

2006. június 3., Pécs

9. Likó István, Borbás Tímea: A flavin-monooxigenáz enzimcsalád homológia

modellezése. MAGYAR BIOKÉMIAI EGYESÜLET 2006. ÉVI

VÁNDORGYŰLÉSE, 2006. aug.30- szept.2., Pécs

10. Borbás Tímea, Jun Zhang, Matt A. Cerny, István Likó, John Cashman: Investigation

of structure and function of a catalitically efficient variant of the human flavin-

containing monooxygenase form 3 (FMO3). 16th INTERNATIONAL

SYMPOSIUM ON MICROSOMES AND DRUG OXIDATIONS, 2006. szept. 3-7.,

Budapest

93

9 ACKNOWLEDGEMENTS In the first place, I would like to thank Professor András Falus for referring me to

Károly Tihanyi in order to do my Ph.D. studies.

I wish to thank my supervisor, Károly Tihanyi for tutoring me. His outstanding

knowledge on drug discovery and his way of thinking about scientific issues made a

strong impact on my work.

I am grateful to Éva Szökő, Szabolcs Szeberényi and Balázs Dalmadi for helping

me with their valuable advices throughout my Ph.D. student years.

I would like to express my sincere gratitude to several people for their

contributions to this thesis: Gyula Beke for synthetizing benzydamine N-oxide

metabolite; Kornél Galgóczy for sharing his experience on setting diabetic conditions;

Viktor Háda for carrying out HPLC-MS measurements; István Likó for helping me with

gene expression determination and also constructing a new model of human FMO3;

Mariann Borsos for providing consultation on data analysis and Teréz Merkl for

excellent technical assistance.

I thank Ottilia Elekes for being a cheerful company and for giving me insights in

systematic, intuitive and successful research work in drug development.

I appreciate a lot my colleagues Bernadett Benkő, Imola Szabó and Erika Czank

to be helpful and kind company and János Leibinger who explained me bioanalytical

methods with great enthusiasm.

I thank Monika Vastag for having me in the In Vitro Metabolism Laboratory and

giving me the possibility to prepare my thesis.

I am grateful to The Hungarian School of Ph.D. Studies, The Hungarian

Pharmaceutical Society Industrial Organization and the Hungarian Pharmaceutical

Chamber for financing my study tour to California and to Prof. John Cashman and Jun

Zhang for the possibility to spend four months in San Diego at the Human

BioMolecular Research Institute allowing me to see the beauty and difficulties of

research.

At last but not at least I thank my parents and my family for their patience, love

and care with which they helped me during these long years.

94