Top Banner
Eur. J. Biochem. 268, 4113–4125 (2001) q FEBS 2001 REVIEW ARTICLE Metabolic conversion as a pre-receptor control mechanism for lipophilic hormones Stefan Nobel 1,2 , Lars Abrahmsen 1 and Udo Oppermann 3 1 Biovitrum AB, Division of Pharmaceuticals, Department of Assay Development and Screening, Stockholm, Sweden; 2 Unit for Biochemical Toxicology, Department of Biochemistry and Biophysics, Wallenberg Laboratory, University of Stockholm,Sweden; 3 Medical Biochemistry and Biophysics, Karolinska Institutet, S-171 77 Stockholm, Sweden The majority of physiological effects mediated by steroids, retinoids and thyroids is accomplished by binding to members of the nuclear receptor superfamily of ligand activated transcription factors. The complex specific effects of lipid hormones depend not only on receptor expression, distribution and interactions, but also on the availability and metabolic conversion of the hormone itself. The cell- specific metabolic activation of inactive hormone precur- sors introduces a further level of hormonal regulation, and constitutes an important concept in endocrinology. The metabolic reactions carried out are achieved by dehydro- genases/reductases, hydroxylases and other enzymes, acting on ligands of the steroid/thyroid/retinoic hormone receptor superfamily. The concept implies that these tissue- and cell- specific metabolic conversions contribute to lipid hormone action, thus pointing to novel targets in drug development. All components of this signalling system, the hormone compounds, the receptor proteins, and modifying enzyme families originate from an early metazoan date, emphasiz- ing the essential nature of all elements for development and diversification of vertebrate life. Keywords: hormone metabolism; hydroxysteroid dehydro- genase, nuclear receptor, steroid, thyroid hormone, retinoid, DHEA, short-chain dehydrogenases/reductases INTRODUCTION Lipid hormones represent chemically distinct classes of molecules mediating a multitude of essential effects in vertebrates and mammals, including control of develop- ment, behaviour, metabolism, reproduction, electrolyte balance, cardiovascular tone, regulation of the immune system and inflammatory response. The main mammalian compounds of importance can be classified as steroids, thyroid hormones and retinoids (Fig. 1). Significant progress in the understanding of the mol- ecular action of lipid hormones has been achieved in the past decades. It has been established that most of their effects are achieved through binding to intracellular recep- tors [1–5], although, several important responses are achieved through binding to plasma membrane components [6]. The characterization of members of the nuclear receptor superfamily and their associated cellular coregu- lators [1–5,7–11] has resulted in a wealth of information regarding how lipid hormones accomplish their specific cellular functions. Association of the receptor with members of the same superfamily, i.e. heterodimerization, and interaction with other transcription factors, such as NF-kB, AP-1 or members of the STAT family, add another dimension of molecular operations, contributing to the tissue-specific effects of this group of hormones [12–17]. Furthermore, work on hormone-response elements (HREs; cognate regulatory DNA sequences) has contributed to a similar extent to our current understanding of the mechan- ism of action [1,2]. A further level of regulation of transcriptional activity is added by interactions with other signal transduction systems, e.g. peptide growth factors, through existence of phosphorylation sites in nuclear receptors [18,19]. It was recognized more than four decades ago that steroid hormones exist in active or inactive forms and that these can be enzymatically interconverted, exemplified by the nonreceptor binding compound estrone and the active hor- mone estradiol, binding to the estrogen receptor (Fig. 1). Primarily regarded as an effect of catabolic reactions, the corresponding inactive metabolites were later considered as precursors, which can be activated in a tissue-specific manner. This concept was established for androgens and estrogens, and coined intracrinology [20–22]. Clearly, this Correspondence to U. Oppermann, Medical Biochemistry and Biophysics, Karolinska Institutet, S-171 77 Stockholm, Sweden. Tel.: 1 46 87287680, E-mail: [email protected] Abbreviations: HSD, hydroxysteroid dehydrogenase; AKR, aldo-keto reductase; SDR, short-chain dehydrogenases/reductases; HRE, hormone response element; GR, glucocorticoid receptor; MR, mineralocorticoid receptor; CYP, cytochrome P450; DHEA, dehydroepiandrosterone; AME, apparent mineralocorticoid excess; SULT, sulfotransferase; GABA, g-aminobutyric acid; VDR, vitamin D receptor; RXR, retinoid receptor X; RA, retinoic acid; CRBP, cellular retinol binding protein; CRABP, cellular retinoic acid binding protein; ALDH, aldehyde dehydrogenase; ADH, alcohol dehydrogenase; SDR, short-chain dehydrogenase/reductase; T3, 3,5,3 0 -tri-iodothyronine; T4 or thyroxine, 3,5,3 0 ,5 0 -tetra-iodothyronine; PGG2, prostaglandin G2; COX, cycloxygenase. (Received 6 March 2001, revised 3 May 2001, accepted 18 June 2001)
13

Metabolic conversion as a pre-receptor control mechanism for lipophilic hormones

Jan 31, 2023

Download

Documents

Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Metabolic conversion as a pre-receptor control mechanism for lipophilic hormones

Eur. J. Biochem. 268, 4113±4125 (2001) q FEBS 2001

R E V I E W A R T I C L E

Metabolic conversion as a pre-receptor control mechanismfor lipophilic hormones

Stefan Nobel1,2, Lars Abrahmsen1 and Udo Oppermann3

1Biovitrum AB, Division of Pharmaceuticals, Department of Assay Development and Screening, Stockholm, Sweden;2Unit for Biochemical Toxicology, Department of Biochemistry and Biophysics, Wallenberg Laboratory, University of Stockholm, Sweden;3Medical Biochemistry and Biophysics, Karolinska Institutet, S-171 77 Stockholm, Sweden

The majority of physiological effects mediated by steroids,

retinoids and thyroids is accomplished by binding to

members of the nuclear receptor superfamily of ligand

activated transcription factors. The complex specific effects

of lipid hormones depend not only on receptor expression,

distribution and interactions, but also on the availability and

metabolic conversion of the hormone itself. The cell-

specific metabolic activation of inactive hormone precur-

sors introduces a further level of hormonal regulation, and

constitutes an important concept in endocrinology. The

metabolic reactions carried out are achieved by dehydro-

genases/reductases, hydroxylases and other enzymes, acting

on ligands of the steroid/thyroid/retinoic hormone receptor

superfamily. The concept implies that these tissue- and cell-

specific metabolic conversions contribute to lipid hormone

action, thus pointing to novel targets in drug development.

All components of this signalling system, the hormone

compounds, the receptor proteins, and modifying enzyme

families originate from an early metazoan date, emphasiz-

ing the essential nature of all elements for development and

diversification of vertebrate life.

Keywords: hormone metabolism; hydroxysteroid dehydro-

genase, nuclear receptor, steroid, thyroid hormone, retinoid,

DHEA, short-chain dehydrogenases/reductases

I N T R O D U C T I O N

Lipid hormones represent chemically distinct classes ofmolecules mediating a multitude of essential effects invertebrates and mammals, including control of develop-ment, behaviour, metabolism, reproduction, electrolytebalance, cardiovascular tone, regulation of the immunesystem and inflammatory response. The main mammaliancompounds of importance can be classified as steroids,thyroid hormones and retinoids (Fig. 1).

Significant progress in the understanding of the mol-ecular action of lipid hormones has been achieved in thepast decades. It has been established that most of their

effects are achieved through binding to intracellular recep-tors [1±5], although, several important responses areachieved through binding to plasma membrane components[6]. The characterization of members of the nuclearreceptor superfamily and their associated cellular coregu-lators [1±5,7±11] has resulted in a wealth of informationregarding how lipid hormones accomplish their specificcellular functions. Association of the receptor withmembers of the same superfamily, i.e. heterodimerization,and interaction with other transcription factors, such asNF-kB, AP-1 or members of the STAT family, add anotherdimension of molecular operations, contributing to thetissue-specific effects of this group of hormones [12±17].Furthermore, work on hormone-response elements (HREs;cognate regulatory DNA sequences) has contributed to asimilar extent to our current understanding of the mechan-ism of action [1,2]. A further level of regulation oftranscriptional activity is added by interactions with othersignal transduction systems, e.g. peptide growth factors,through existence of phosphorylation sites in nuclearreceptors [18,19].

It was recognized more than four decades ago that steroidhormones exist in active or inactive forms and that thesecan be enzymatically interconverted, exemplified by thenonreceptor binding compound estrone and the active hor-mone estradiol, binding to the estrogen receptor (Fig. 1).Primarily regarded as an effect of catabolic reactions, thecorresponding inactive metabolites were later considered asprecursors, which can be activated in a tissue-specificmanner. This concept was established for androgens andestrogens, and coined intracrinology [20±22]. Clearly, this

Correspondence to U. Oppermann, Medical Biochemistry and

Biophysics, Karolinska Institutet, S-171 77 Stockholm, Sweden.

Tel.: 1 46 87287680, E-mail: [email protected]

Abbreviations: HSD, hydroxysteroid dehydrogenase; AKR, aldo-keto

reductase; SDR, short-chain dehydrogenases/reductases; HRE,

hormone response element; GR, glucocorticoid receptor; MR,

mineralocorticoid receptor; CYP, cytochrome P450; DHEA,

dehydroepiandrosterone; AME, apparent mineralocorticoid excess;

SULT, sulfotransferase; GABA, g-aminobutyric acid; VDR, vitamin D

receptor; RXR, retinoid receptor X; RA, retinoic acid; CRBP, cellular

retinol binding protein; CRABP, cellular retinoic acid binding protein;

ALDH, aldehyde dehydrogenase; ADH, alcohol dehydrogenase; SDR,

short-chain dehydrogenase/reductase; T3, 3,5,3 0-tri-iodothyronine; T4

or thyroxine, 3,5,3 0,5 0-tetra-iodothyronine; PGG2, prostaglandin G2;

COX, cycloxygenase.

(Received 6 March 2001, revised 3 May 2001, accepted 18 June 2001)

Page 2: Metabolic conversion as a pre-receptor control mechanism for lipophilic hormones

idea adds a further mechanistic principle to hormonephysiology, and implies that metabolic conversion comple-ments the cascade of ligand-binding to the receptor andassociated regulatory transcriptional events, as outlined inFig. 2.

The aim of this review is to demonstrate by selectedexamples, that enzymatic regulation of active hormones,constitutes a prereceptor control (`switch') mechanism,which is a common theme along the different classes oflipid hormones, and which is not restricted to specificsteroid hormones. The majority of these prereceptorregulatory enzymes are NAD(P)(H) dependent hydroxy-steroid dehydrogenases (HSDs). At present, all HSDsprincipally involved in hormone metabolism belong toeither the short-chain dehydrogenase/reductase (SDR)superfamily [23,24], or to the aldo-keto reductase (AKR)superfamily [25]. Additional enzymes involved in hormonemetabolism belong to different classes of oxidoreductases,e.g. cytochrome P450 (CYP) enzymes and different con-jugating phase II enzymes [26]. However, these reactionsmostly constitute irreversible transformations involved insynthetic or catabolic pathways.

H O R M O N E S Y N T H E S I S

There are important differences in the synthetic routes ofthe various lipophilic hormones and vitamins. Chemically,the different steroid hormones, including vitamin D, andretinoids are isoprene condensation products, whereasthyroids are derivatives of the amino acid tyrosine.

In mammals, the location of steroid hormone synthesisdepends on the class of hormone. These production sites aremainly the `classical' endocrine glands, such as adrenal,gonads, uterus and placenta, although important species andgender differences exist. However, auto- or paracrinesynthesis of steroids in `nonclassical endocrine' organs,e.g. brain, heart or thymus, has been reported [27±29].Furthermore, tissue-specific expression of critical keymetabolic enzymes creates the possibility to regulatelocally the amount of available hormone from circulatingprecursors.

The different biosynthetic routes leading to the definedclasses of steroid hormones are performed predominantlyby distinct oxidases and oxidoreductases of the cytochromeP450 and SDR superfamilies (reviewed in [26]). Different

Fig. 1. Lipid hormones and some metabolic conversion routes as discussed in this text. Receptor ligands of different hormone classes are

highlighted by grey boxes. Double arrows indicate reversible steps, carried out by different isozymes or different enzyme classes, whereas bold,

single-headed arrows point to `irreversible' reactions. Enzymes are indicated by numbers as follows: 1, steroid sulfatase; 2, steroid sulfotransferase;

3, 3b-hydroxysteroid dehydrogenase/D4,5 isomerase; 4, 17b-hydroxysteroid dehydrogenase; 5, 5a-reductase; 6, 3a-hydroxysteroid dehydrogenase;

7, 11b-hydroxysteroid dehydrogenase; 8, 20a-hydroxysteroid dehydrogenase; 9, retinaldehyde dehydrogenase; 10, retinol dehydrogenase

4114 S. Nobel et al. (Eur. J. Biochem. 268) q FEBS 2001

Page 3: Metabolic conversion as a pre-receptor control mechanism for lipophilic hormones

inborn errors of metabolism, caused by mutations insynthetic enzymes are now well established, emphasizingthe critical role of steroids in mammalian physiology anddevelopment [26,30,31]. Of further importance, in humansand other primates, is the extra-gonadal and extra-placentalexpression of key enzymes such as steroid sulfatase,3bHSD/D4,5 isomerase, 17b-HSD isozymes and CYP19(aromatase), e.g. in breast tissue, to locally produce bothandrogen and estrogen hormones [20±22] from theadrenal precursors dehydroepiandrosterone (DHEA) andits sulfonate ester dehydroepiandrosterone sulfate(DHEA-S).

S T E R O I D H O R M O N E M E T A B O L I S M

Only steroids displaying distinct structural properties act asligands by binding to the different nuclear receptors. Forexample, a nearly planar steroid ring structure and a 3-oxoconfiguration are common denominators of glucocorticoid,mineralocorticoid, progesterone, estrogen and androgenhormones for succesful association with their respectivereceptors. Metabolic transformations of these chemicalconfigurations therefore have a profound effect on receptorbinding, e.g. loss of the planar structure by 5b reduction ofdouble bonds at C5 results in complete loss of receptor

association of steroid hormones, while reduction of the3-oxo group to 3a or 3b hydroxyls leads to reduced bindingproperties. There are further specific reactions for inactivat-ing active steroid hormones. Besides reduction of doublebonds these comprise hydroxylation, subsequent conjuga-tion (e.g. with sulfate or glucuronic acid) and furtheroxidoreductive reactions. Reduction of C4±5 or C5±6double bonds to the 5a or 5b isomers are virtually irre-versible reactions in mammals, while many other reactionsmust be regarded as `biologically reversible'. Tissue-specific expression of enzymes catalyzing a reversion ofreaction leads to an intricate balance or `shuttle', allowingspecific regulation of steroid ligand availability. Several ofthese pairs of `switch' mechanisms are discussed in thefollowing sections.

G L U C O C O R T I C O I D S A N DM I N E R A L O C O R T I C O I D S

In humans, the adrenal glands secrete three different classesof steroids; glucocorticoids, mineralocorticoids and andro-gens (mainly the sex steroid precursors DHEA and itssulfonated derivative DHEA-S, see below). The glucocorti-coids (cortisol in man and most mammals, corticosteronein rodents and lower vertebrates) exert potent effects oncellular function in essentially all organ systems. Theeffects include regulation of carbohydrate and amino-acidmetabolism, maintenance of blood pressure, modulation ofthe stress and inflammatory responses [32], maturation offetal organ systems and metabolic adaptation during preg-nancy [33]. In contrast to this broad range of effects,mineralocorticoids (in humans mainly aldosterone) primar-ily affect the extracellular balance of sodium and potassiumin target tissues such as kidney, colon and salivary glands[34,35].

The hydroxyl group at C11 is critical for cortisol to exertits biological effect, i.e. is necessary for cortisol to bindto the glucocorticoid receptor (GR). The GR and themineralocorticoid receptor (MR) share 57% and 96%amino-acid identities in steroid and DNA binding domains,respectively [36,37]. The mineralocorticoid hormone aldo-sterone binds only to MR with high affinity (Kd 1 nm). Incontrast, cortisol/corticosterone have nanomolar affinity forboth receptors and interact stronger with MR (Kd 1 nmcompared to 10 nm for GR). Despite this high affinity,cortisol/corticosterone do not bind to the MR in mineralo-corticoid target tissues (e.g. kidney, colon, salivary glands).The reason is local inactivation by oxidation of the C11hydroxyl group (dehydrogenation), catalyzed by the enzyme11b-hydroxysteroid dehydrogenase 2 (11b-HSD2). Theimportant function of 11b-HSD2 to protect MR frominappropriate activation by cortisol/corticosterone is clearlyillustrated in the syndrome of apparent mineralocorticoidexcess (AME), associated with hypertension and hypokal-emia caused by inactivating mutations in the 11b-HSD2gene [[38,39]. As expected, mice with a deleted 11b-HSD2gene develop hypertension [40]. However, the role of11b-HSD2 is not exclusively to protect MR, demonstratedby its important role in the placenta protecting the fetusagainst high levels of maternal glucocorticoids [33]. Somenonepithelial tissues have MR but lack 11b-HSD2. Here,not fully characterized mechanisms lead to different effectsfrom glucocorticoid or aldosterone binding to the MR [41].

Fig. 2. Principle of local lipophilic hormone regulation (steroids,

retinoids and thyroid hormones). Model of regulation of local

hormone concentration, representing a `shuttle' system between active

hormone (HA) and inactive hormone (HI), mediated by tissue-specific

metabolizing enzymes (activating, Ea, and inactivating, Ei). The

occurrence of these enzymes determines the hormonal responsiveness

of the particular cell. Note that opposing enzymes may not be localized

within the same cell. Plasma binding proteins exist for different types

of ligands, e.g. CBG (corticosteroid binding globulin) selectively

binding cortisol, thereby influencing the level of `free' hormone, which

can enter the cell. NR � nuclear receptor.

q FEBS 2001 Enzymatic control of lipid hormones (Eur. J. Biochem. 268) 4115

Page 4: Metabolic conversion as a pre-receptor control mechanism for lipophilic hormones

In the hippocampus, MR and GR coexist and monitorcortisol/corticosterone levels; due to the higher affinity, theformer is occupied at basal levels of glucocorticoid, whilethe latter is only occupied at peak levels [42].

Peripheral formation of glucocorticoids has gained con-siderable attention over the last few years, as activation ofcortisone by 11-oxo reduction to cortisol has been shown tooccur in several peripheral tissues (reviewed in [43]). Thisreaction is catalyzed by 11b-hydroxysteroid dehydrogenasetype 1 (11b-HSD1), initially believed to be the inactivatingenzyme involved in AME (above), but later shown tofunction as an 11-oxo-reductase in vivo [44]. This isoformis coexpressed in most tissues with GR (primarily liver,omental fat, gonads, brain, skeletal muscle and vasculature[45]). Gene-deletion experiments in mouse indicate that thisenzyme is important both for maintenance of normal serumcortisol levels, and for upregulation of key hepatic gluco-neogenic enzymes [46]. Clinical observations of deficiencyin cortisone activation have been made [47,48], although sofar no clear evidence for a primary deficiency of 11b-HSD1has been reported [49,50]. However, there are severalclinical indications for a role of 11b-HSD1 in maintainingcortisol action, and therefore it has been postulated that11b-HSD1 has a role in the etiology of insulin resistance[51] and control of insulin secretion [52].

Thus, to date two enzymes have been identified whichmediate interconversion between cortisol and cortisone,each enzyme irreversible in vivo and together yielding`biological reversibility'. The ratio of cortisol to cortisonein circulation is determined by the balance between theactivities of those two enzymes, mainly 11b-HSD1 in liverand 11b-HSD2 in kidney. Yet, the important principle isthat the corticosteroid action upon individual target cells isdetermined by the enzyme action within the cells and notby the circulating steroid levels alone.

D H E A A N D D H E A - S

In humans, sex steroids are synthesized in peripheral orgonadal tissues: a substantial amount (up to 30±50% oftotal androgens in males) are produced by extragonadalconversion of inactive precursors, and in females peripheralestrogen formation might be even more important [21]. Theprecursor steroids DHEA and to a lower extent pregneno-lone, and its sulfonated derivatives DHEA-S and pregne-nolone-S are secreted from the adrenals in large amounts, aunique feature for humans and other higher primates. Asubstantial reduction (70±95%) in the formation of DHEAand DHEA-S occurs during ageing in parallel to a reductionof sex steroid levels [53]. This is the background to promoteDHEA as a youth drug, although at present the quantitativeimportance of plasma DHEA as a precursor for testosteroneand estradiol is poorly understood. DHEA-S is consideredas an inactive reservoir of DHEA, which is locally con-verted by enzymatic removal of the sulfonate prior tofurther conversion to steroid hormones [54]. Furthermore,kinetic parameters such as binding of sulfoconjugates toplasma proteins, hepatic and urinary clearance, prolongserum half-life by two orders of magnitude [55]. Severalenzymes of the sulfotransferase (SULT) and sulfatasefamilies are involved in either addition or removal of thesulfonate group, together forming another example ofbiological reversibility or `switch' [56±58]. Sulfonation of

steroids may also function as a trapping mechanism for theintracellular storage of steroid hormone or precursors, e.g.in steroidogenic tissues such as the adrenal cortex wheresulfotransferases are highly expressed [59].

A N D R O G E N S A N D E S T R O G E N S

Active androgens and estrogens are in the 17b-hydroxyconfiguration, whereas the 17-oxo derivatives are notbinding to androgen and estrogen receptors, respectively.The conversion between the inactive and the active forms(estradiol/estrone and testosterone/androstenedione) is cata-lyzed by differentially expressed isozymes belonging to the17b-hydroxysteroid dehydrogenase (17b-HSD) family. Spa-tial and temporal activation and deactivation, is achieved byisozymes that display distinct reaction directions in vivo.Mostly, gonadal tissues convert 17-oxo steroids to17-hydroxy steroids, while the opposite holds true formany extragonadal tissues. The different isoforms within agiven species share less than 30% identity, while theorthologs are 70% or more identical, facilitating identifica-tion and consistent nomenclature in different species. Todate, 11 different 17b-HSD isoenzymes have been charac-terized (Table 1). However, the identification of the in vivo-relevant reaction is not always straightforward, as most17b-HSDs are able to catalyze the reversible reactionsin vitro [60]. Furthermore, several 17b-HSDs display broadsubstrate specificities, e.g. with progestins, retinoids, xeno-biotics and fatty acids, which makes the identification ofthe in vivo role a challenging task [61], further exemplifiedby 17b-HSD isoforms 4 and 9 [60] (see below). Phylo-genetic analysis indicates that enzymes regulating accessof active estrogens and androgens (i.e. 17b-HSD2 and17b-HSD6) are related to retinoid converting enzymes inC. elegans [62], the latter displaying 65% sequence identityto retinol dehydrogenase type 1.

Fig. 3. Steroid conversions involved in local estrogen activation in

breast tissue. The enzymes involved are indicated by numbers: 1,

steroid sulfatase; 2, steroid sulfotransferase; 3, 3b-HSD/ketosteroid

isomerase; 4, aromatase (CYP 19); 5. Estrone sulfotransferase; 6,

estrogen sulfatase; 7, 17b-HSD1; 8. 17b-HSD2 and 17b-HSD4.

DHEA-S, estrone-S and estradiol-S indicate the sulfonated derivatives.

4116 S. Nobel et al. (Eur. J. Biochem. 268) q FEBS 2001

Page 5: Metabolic conversion as a pre-receptor control mechanism for lipophilic hormones

Ta

ble

1.

En

zym

esin

vo

lved

inti

ssu

e-sp

ecif

ich

orm

on

e`s

wit

ch'

mec

han

ism

s.D

etai

led

info

rmat

ion

on

AK

Ren

zym

esis

avai

lable

at:

htt

p:/

/ww

w.m

ed.u

pen

n.e

du/a

kr.

Abbre

via

tions

use

d:

AK

R:

aldo-k

eto

red

uct

ases

;S

DR

:sh

ort

-ch

ain

deh

yd

rog

enas

es/r

edu

ctas

es;

MD

R:

med

ium

-chai

ndeh

ydro

gen

ase/

reduct

ase.

DD

H:

dih

ydro

dio

ldeh

ydro

gen

ase;

PG

DH

:pro

stag

landin

deh

ydro

gen

ase;

h:

hum

an;

r:ro

den

t;G

C:

glu

coco

rtic

oid

;S

UL

T:

sulf

otr

ansf

eras

es;

nc:

no

men

clat

ure

;50 D

:50 d

eiodin

ases

;A

LD

H:

aldeh

yde

deh

ydro

gen

ases

;D

H:

deh

ydro

gen

ase;

CR

:ca

rbonyl

reduct

ase,

DH

T:

dih

ydro

test

ost

erone.

Num

ber

ing

inco

lum

n

En

zym

efo

rm/c

lass

indic

ates

dis

tin

ctp

rote

ins

wit

hover

lap

pin

gac

tivit

ies

or

spec

ific

itie

s

Act

ivit

y

En

zym

e

form

/cla

ss

Pro

tein

Fam

ily

Inv

ivo

reac

tion

Mai

n

Funct

ion

Oth

ersu

bst

rate

spec

ific

itie

sS

pec

ies

Ref

.

3a

-HS

D1

AK

Rre

duct

ase

andro

gen

inac

tivat

ion

DD

H,

PG

DH

,h/r

[69]

2A

KR

reduct

ase

5a

-DH

Tin

acti

vat

ion

iden

tica

lto

17b

-HS

D5

h/r

[69]

3A

KR

deh

ydro

gen

ase

5a

-DH

Tpro

duct

ion

±h/r

[69]

3b

-HS

D1

SD

Rdeh

ydro

gen

ase/

isom

eras

est

eroid

synth

esis

±h/r

[134]

2S

DR

deh

ydro

gen

ase/

isom

eras

est

eroid

synth

esis

±h/r

[134]

6S

DR

deh

ydro

gen

ase/

isom

eras

est

eroid

synth

esis

±r

[135]

11b

-HS

D1

SD

Rre

duct

ase

GC

acti

vat

ion

mic

roso

mal

CR

h/r

[41,1

36]

2S

DR

deh

ydro

gen

ase

GC

inac

tivat

ion

±h/r

[41]

17b

-HS

D1

SD

Rre

duct

ase

estr

adio

lpro

duct

ion

20a

-HS

Dh/r

[58]

2S

DR

deh

ydro

gen

ase

sex

ster

oid

inac

tivat

ion

20a

-HS

Dh/r

[58]

3S

DR

reduct

ase

test

ost

erone

pro

duct

ion

h/r

[58]

4S

DR

deh

ydro

gen

ase

estr

adio

lin

acti

vat

ion

b-o

xid

atio

nh/r

[58]

5A

KR

reduct

ase

test

ost

erone

pro

duct

ion

3a

-HS

D,

DD

Hh/r

[58,6

9]

6S

DR

deh

ydro

gen

ase

DH

Tin

acti

vat

ion

3a

-HS

Dh/r

[58]

7S

DR

reduct

ase

estr

adio

lpro

duct

ion

h/r

[58]

8S

DR

deh

ydro

gen

ase

estr

adio

lin

acti

vat

ion

h/r

[58]

9S

DR

sex

ster

oid

inac

tivat

ion

3a

-HS

D,

reti

nol

DH

r[1

01]

10

SD

Rdeh

ydro

gen

ase?

sex

ster

oid

inac

tivat

ion?

b-o

xid

atio

n,

3a

-HS

Dh/r

[71]

11

SD

Rdeh

ydro

gen

ase?

sex

ster

oid

inac

tivat

ion?

h[1

37]

20a

-HS

Dn

cn

ot

avai

lab

leA

KR

reduct

ase

pro

ges

tero

ne

inac

tivat

ion

DD

H,

aldose

reduct

ase

h/r

[78,1

18]

Ste

roid

sulf

atas

esu

lfat

ases

sulf

ate

este

rhydro

lysi

san

dro

gen

pre

curs

or

acti

vat

ion

?h/r

[56]

Ste

roid

sulf

otr

ansf

eras

eS

UL

Tsu

lfonyl

tran

sfer

sulf

onyl

tran

sfer

,in

acti

vat

ion

±h/r

[54]

Ret

ino

ld

ehy

dro

gen

ase

9S

DR

/MD

Rdeh

ydro

gen

ase

reti

noic

acid

synth

esis

oft

en3a

-HS

Dh/r

[93]

Ret

inal

deh

yd

ed

ehy

dro

gen

ase

3A

LD

Hdeh

ydro

gen

ase

reti

noid

acid

synth

esis

±h/r

[102]

Iod

oth

yro

nin

ed

eidoin

ase

150 D

sulf

ated

thyro

nin

edei

odin

atio

nth

yro

idhorm

one

inac

tivat

ion

±h/r

[106]

250 D

T4-T

3th

yro

idhorm

one

acti

vat

ion

±h/r

[105]

350 D

T4,

T3-

rT4,

rT3

thyro

idhorm

one

inac

tivat

ion

±h/r

[107]

q FEBS 2001 Enzymatic control of lipid hormones (Eur. J. Biochem. 268) 4117

Page 6: Metabolic conversion as a pre-receptor control mechanism for lipophilic hormones

The role of temporally and spatially differentiatedexpression is demonstrated by the estrous cycle in rodents,during which at least four different 17b-HSD isozymes areexpressed. 17b-HSD1 is found in granulosa cells ofdeveloping follicles, activating estrone to estradiol. Thereductive 17b-HSD2 is expressed in epithelial endometrialcells. The type 7 isoform is responsible for local estradiolbiosynthesis in luteinized cells, while 17b-HSD8 sup-posedly has an inactivating role in the cumulus cells of theovary [22,63].

Furthermore, 17b-HSD1 catalyzes formation of estradiolfrom estrone in breast tissue. This local estradiol productionis important for tumor growth, making the enzyme anattractive target for inhibitor development [20]. A similarrole has been suggested for 17b-HSD4 in colonic cancer,where malignant biopsy material displayed an alteredactivity of this isozyme [64]. Local estrogen formationcan also be reduced by blocking the CYP19 (cytochromeP450 aromatase) mediated production of estrone fromandrostenedione [65]. Importantly, estradiol can be inacti-vated in a tissue-specific manner by sulfonation viaestrogen sulfotransferase catalyzed transfer of a sulfonateradical to the 3-hydroxy group. A mechanistic principle wassuggested by recent data, implicating that the estrogenactivity of hydroxylated polychlorinated biphenyls is due toinhibition of estrogen sulfotransferase, resulting in increasedestrogenic activity [66]. The sulfonate group can beenzymatically removed by estrogen sulfatase, providinganother mechanism to regulate local estradiol concentration[57]. Further metabolic alterations, e.g. change in redoxstate have a profound effect on steroid profiles and clinicalimplications have been suggested [67±69]. Figure 3 sum-marizes the reaction pathways centering around estradiol,e.g. in breast tissue and illustrates the importance ofisozymes in control of active hormone level.

The intracrine formation of active androgens in prostateand testes is essential for development of the male genitaltract [22]. 17b-HSD3, converting androstenedione intotestosterone, is almost exclusively expressed in testes, anddeficiency during fetal development leads to male pseudo-hermaphroditism [31]. Another genetic defect leading tomale pseudohermaphroditism is deficiency in steroid5a-reductase type 2, catalyzing the formation of theessential androgen dihydrotestosterone [70]. Due to thisisoform the ratio between testosterone and dihydrotesto-sterone is 1 : 10 in prostate, in contrast to the ratio of 10 : 1in plasma. However, 5a-reductase inhibitors, of use to treatboth benign prostate hyperplasia and prostate cancer,should display overlapping inhibition characteristics againsttype 1 and 2 isozymes. This is demonstrated by the widelyused inhibitor finasteride, which only modestly decreasesclinical signs, likely due to lack of type 1 5a-reductaseinhibition [70]. The situation is rendered more complexthrough metabolic pathways involving 3a-HSD isozymes,mostly belonging to the AKR superfamily, with a distinctand specific expression and isozyme pattern, leading toan androgen shuttle between active and inactive forms[71±73].

P R O G E S T I N S

The progestin progesterone is required to support gestationin humans, and is synthesized initially by the corpus luteum

and during term by the placenta. Chemical abortion, such astreatment with the progesterone antagonist mifepristine(RU486) prevents ovulation by inducing luteolysis andpremature menstruation. In many but not all species,progesterone is essential for maintaining pregnancy, andaccordingly, metabolism to the weaker progestin 20a-hydroxy progesterone by 20a-HSD can be associated withtermination of luteal and placental stage pregnancy[74±77], indicating that 20a-HSD works as a progestin`switch' in a species-specific manner. In testis or theadrenals, the major substrates for 20a-HSD may be17a-hydroxypregnenolone or 17a-hydroxy progesterone,leading to 20a-OH steroids. These are not substrates for theCYP450 17,20-lyase and thus cannot be converted intoprecursors of androgens and estrogens, in contrast to theparent steroids. A novel role for progesterone metabolizingenzymes, such as 20a-HSD, has been postulated in theprotection of MR against high levels of progesterone [78],similar in concept as the role of 11b-HSD2 in the MRprotection against cortisol occupancy.

Several 20a-HSD isoforms have been identified thus far[60]. The ovarian 20a-HSDs characterized display high-sequence similarity to cytosolic rat liver 3a-HSD, and thusare members of the AKR superfamily [79]. Again, theoverlapping substrate specificities of 17b-HSD1 and17bHSD2 suggest roles in 20a-OH metabolism ofprogestins. 17b-HSD1 inactivates progesterone while17b-HSD2 converts 20a-hydroxyprogesterone to proges-terone, contributing to the estrogen- and progestin-dependentregulation of endometrial growth [60]. However, it isanticipated that further forms exist [76,80].

N E U R O S T E R O I D S

The concept of local activation and inactivation reactionstakes another twist in the brain, where steroids havemodulating functions in addition to their role as ligands ofnuclear receptors (reviewed in [81]). It has long beenknown that steroids have multiple effects on nerve cells.Initially, it was postulated that these steroids weresynthesized peripherally by endocrine glands, until theirsynthesis in the brain was detected by Baulieu andcoworkers [82]. The term `neurosteroids' has been coinedto designate steroids synthesized from cholesterol or otherprecursors in the nervous system. Neurosteroids allosteri-cally modulate the affinity of receptors of several types,including N-methyl-d-aspartate receptors, nicotinic recep-tors and g-aminobutyric acid (GABA)A receptors. Forexample nanomolar concentrations of tetrahydrosteroidsincrease the affinity of the GABAA receptor for GABA[83]. These tetrahydrosteroids are locally formed from5a-reduced steroids (e.g. 5a,3a-tetrahydroprogesteronefrom 5a-dihydroprogesterone) by 3a-HSD. This is sup-ported by recent results, indicating that changes in thesensitivity of GABAA receptor are associated with fluctu-ations in endogenous levels of progesterone and 5a-dihydroprogesterone, occuring during the menstrual cycleand pregnancy [84]. This effect may be responsible forsome of the pre±menstrual syndrome symptoms [84]. Inaddition to the modulation of GABAA affinity by tetra-hydrosteroids, a decreased affinity of GABAA receptors bysulfoconjugates of pregnenolone and DHEA has beenreported [85]. However, signalling through other receptor

4118 S. Nobel et al. (Eur. J. Biochem. 268) q FEBS 2001

Page 7: Metabolic conversion as a pre-receptor control mechanism for lipophilic hormones

types, e.g. the oxytocin receptor, is also modulated bypregnane derivatives [86] again demonstrating the complexinterplay between peptide and steroid ligands. An addi-tional role of neurosteroids in regulation of tubulin poly-merization was recently reported [87]. In summary, steroidslocally produced in brain are involved in the control ofmetabolic, behavioral, and psychical processes includingcognition, stress, anxiety and sleep [83,85].

V I T A M I N D

Vitamin D was identified as the antirachitic principle inrickets. It plays a central role in calcium and phosphatehomeostasis, and is essential for the development andmaintenance of osseous tissue [88]. The isolation of fat-soluble, antirachitic components in, e.g. fish-oil, and thediscovery of photolytic cleavage of 7-dehydrocholesterolby UV light, led to the elucidation of vitamin D synthesisand structures. Several natural, dietary compounds werefound to display biological activity, e.g. ergocalciferol(vitamin D2) and cholecalciferol (vitamin D3). Both areformed by formation of secosteroids through pericyclic UVcleavage of steroid ring B, a reaction occuring in the skin.Further research led to the identification of more potentcompounds. This activation is achieved through 25-hydroxylation of the steroid alkyl-side chain [25-(OH)D3],and further 1a-hydroxylation, to yield 1,25-dihydroxyvitaminD3 [1,25(OH)2,3]. These reactions are cytochrome P450-mediated, predominantly occuring in liver and kidney,respectively [89]. The discovery of the vitamin D receptor(VDR) and its heterodimerization with retinoid receptor X(RXR), the complex interactions with coactivators, andexpression of enzymes involved in vitamin D3 synthesis,has resulted in a detailed picture of vitamin D action [88]. Itis now accepted that vitamin D effects are not onlyrestricted to `classical' target tissues, such as bone, kidneyand intestine, but also play important roles in `nonclassical'target tissues, such as skin, pancreas, muscle, hematopoieticand malignant cells, the nervous system and the immunesystem [88,90,91]. Besides the `classical' regulation ofcalcium and phosphate metabolism [92], these `nonclassical'effects include antiproliferative and prodifferentiatingactions in hematopoietic cells and suppression of theimmune system (similar to glucocorticoids) by shifting theTh1/Th2 thymocyte balance [93±95]. One example of localhormone regulation is the expression of 1a-hydroxylase inactivated macrophages [90] which results in local activationand potentiation of 25(OH)D3 effects, leading to altereddifferentiation and maturation of monocytes.

R E T I N O I D S

For many years, naturally occuring retinoids and theiractive metabolites have been recognized as essential mol-ecules regulating growth, reproduction, vision, resistance toinfection, as well as being associated with schizophrenia[96±99]. Retinoic acid (RA) directs a variety of essentialbiological effects by modulating gene expression duringdevelopment and postnatally, to control differentiation orcell-death of numerous cell types in several organs [96,98].Retinoids are stored as retinyl esters, and transported byseveral classes of binding proteins, emphasizing that themajority of cellular retinoids exist as protein-bound

molecules. The binding proteins consist of cellular retinolbinding protein (CRBP) and cellular retinoic acid bindingprotein (CRABP), which together with the retinoid forms atight hormone-binding protein `cassette' [98,100]. Afterhydrolysis of esterified forms, retinol (vitamin A) under-goes metabolic activation by dehydrogenation into retinal(carried out by retinol dehydrogenases), followed by irre-versible oxidation into the active hormone RA (performedby retinaldehyde dehydrogenases), a metabolic route resemb-ling the pathway of ethanol oxidation to acetaldehyde andacetic acid in mammals. Indeed, all steps in RA synthesiscan be catalyzed by several isozymes of the aldehydedehydrogenase (ALDH) and alcohol dehydrogenase (ADH/medium chain dehydrogenase reductase) superfamilies[101,102], at least in vitro using unbound retinoids. Inlight of the tight binding to CRBP/CRABP, the enzymesresponsible for retinoid acid metabolism in vivo are con-sidered to metabolize protein-bound retinoids [98],although this point is at present insufficiently understood.CRBP deficient mice show normal retinoid metabolism[103], whereas deletion of class I and IV ADH in micesuggest a role of these isozymes in retinol metabolism[104,105]. Several further cytosolic and microsomal retinoldehydrogenases have now been identified to catalyze theCRBP-bound retinol dehydrogenase reaction [98], thusconstituting an important regulatory mechanism. At least 10different cDNA clones from rodent and human species havebeen isolated, all of which encode microsomal members ofthe short-chain dehydrogenase/reductase (SDR) superfamily[23,98,106], and which catalyze the conversion of all-transretinol or various cis-retinol isomers into the correspondingretinals. Notably, most of these enzymes display multiple, tosome extent species-dependent substrate specificities, e.g. as3a-hydroxysteroid dehydrogenases in steroid hormone meta-bolism [98,107±109], thereby adding a further link betweensteroid and retinoid hormone pathways (see Table 1).

The two major aldehyde dehydrogenases involved in theirreversible oxidation to the retinoic acids are members ofthe aldehyde dehydrogenase superfamily [110]. Furtherreactions, mainly leading to inactive metabolites are carriedout by different enzymes of the cytochrome P450 super-family [98]. Some purified enzymes of this family catalyzethe conversion of retinol into retinal (similar to some CYPmediated hydroxysteroid dehydrogenase activities) and ofretinal into retinoic acid, but with unfavourable kineticconstants, suggesting that these CYP mediated reactions donot significantly, if at all, contribute to a retinoid `shuttle'metabolism in vivo.

T H Y R O I D H O R M O N E S

Thyroid hormones regulate essential functions duringdevelopment and normal life in most organs. The majorproducts secreted from the thyroid gland are 3,5,3 0-tri-iodothyronine (T3) and 3,5,3 0,5 0-tetra-iodothyronine (T4;thyroxine). Thyroxine is < 8- to 10-fold more abundant inthe thyroidal secretion, but the thyroid hormone receptorbinds T3 with a 10-fold higher affinity compared to T4.Approximately 85% of the secreted T4 is metabolizedperipherally to the active hormone T3 and the inactivecompound `reverse T3' (3,3 0,5 0-tri-iodo-thyronine, rT3),catalyzed by iodothyronine 5 0deiodinase (5 0D) isozymes(see Table 1, Fig. 1) [111]. These enzymes belong to a

q FEBS 2001 Enzymatic control of lipid hormones (Eur. J. Biochem. 268) 4119

Page 8: Metabolic conversion as a pre-receptor control mechanism for lipophilic hormones

recently identified family of eukaryotic selenoproteins[112±115], which thus far comprises three distinctisozymes. These enzymes display a spatial and temporaltissue-specific expression pattern and consequently, areinvolved in local activation or inactivation of this class oflipophilic hormones. Type I deiodinase (5 0D-I) primarilyinactivates sulfated thyronines, whereas type II deiodinase(5-D-II) catalyzes the activation step from T4 to T3, and ishighly expressed in fetal and adult brain, heart, skeletalmuscle and placenta. Interestingly, 5-D-II activity appa-rently is important to maintain adequate T3 levels not onlyunder basal conditions, but also under limiting amounts ofT4, as often observed under iodine deficiencies [112,116].Type 3 deiodinase (5 0-D-III) is considered to be a hormoneinactivating enzyme as it converts T4 and T3 to theirinactive metabolites (reverse T3 and 3,3 0-di-iodothyronine,respectively). Thus, the expression pattern of these iso-zymes in addition to other thyronine modifying enzymes(e.g. sulfatases and sulfotransferases) dictates to an appre-ciable extent the tissue-specific action of thyroid hormones[112,117].

A R A C H I D O N I C A C I D D E R I V A T I V E S

Eicosanoids, derived from arachidonic acid, are a group oflipid mediators distinct from the lipid hormones bothchemically and by the signal transduction they convey.These short-lived compounds are formed locally and exert aparacrine or autocrine mode of action, mostly mediatedthrough receptors located in the plasma membrane.However, some actions are mediated through members ofthe nuclear receptor family [4] and some metabolic stepsshow a similar pattern as the steroid `shuttle' systems.

Eicosanoids play essential roles in, e.g. inflammatoryprocesses, and form the main distinct classes of prosta-glandins, thromboxanes, leukotrienes and lipoxins [118].Upon release from phospholipids by phospholipases,arachidonic acid is converted into prostaglandin G2(PGG2) by two different types of cyclooxygenases(COX1 and COX2) [118,119]. This intermediate serves asstarting point for the synthesis of all classes of prostaglan-dins (either nonenzymatically or catalyzed through distinctPG synthases) and for the synthesis of thromboxanes(through thromboxane synthase). Leukotrienes are derivedfrom arachidonic acid via 5-lipoxygenase to form theendoperoxide LTA4, which is an intermediate for formationof LTB4 (via LTA4 hydrolase), and glutathione adducts viaglutathione transfer. These pathways involve mostly irre-versible nonenzymatic and enzyme-mediated steps. How-ever, several dehydrogenase and reductase reactions are ofconsiderable importance, mainly in the metabolism ofprostaglandins. Among the multitude of reactions possible,the oxidation at the 15-OH group of prostaglandins leads toinactive metabolites. This reaction is catalyzed by distinctNAD(P)1-dependent prostaglandin dehydrogenases, belong-ing to the SDR superfamily [120,121]. Several attempts todevelop 15-OH prostaglandin dehydrogenase antagonists,useful as antiulcer, antithrombotic or anticancer agents havebeen initiated, however, mainly are discontinued.

Another enzyme of interest, the NADPH-dependentenzyme 11-ketoreductase, is involved in the stereospecifictransformation of PGD2 to the metabolite 9a, 11b-PGF2.In humans, this enzyme has been found mainly in liver and

lung [122]. Interestingly, the 9a, 11b-PGF2 compoundshows a different activity profile, and has been found tomediate contraction of bronchial smooth muscle, coronaryarteries, to inhibit platelet aggregation, to induce natriuresisand to be a vasopressor substance [122].

E V O L U T I O N O F H O R M O N EM E T A B O L I S M A N D R E C E P T O RP A T H W A Y S

As outlined above, signalling via lipid hormones andmediators and its biotransformation processes complementeach other in the common mechanistic principle of generegulation. All components of the systems described havedistinct phylogenetic roots and evolved separately, reflect-ing the transition of xenobiotics to become membranecomponents, ligands for binding proteins, or substrates formetabolizing enzymes [123±125]. This is clearly exempli-fied in the case of steroids. These molecules were probablyfirst formed in a period when sufficient oxygen was presentin the atmosphere, as the synthesis of cholesterol is strictlydependent on the occurrence of oxygen. These cholesterolderivatives might have served as components to regulatemembrane fluidity and then adapted to intracellular signal-ling molecules, either by binding to receptors or byformation of other signalling mechanisms, such as raftlipid microdomains [126]. At the same time, membranesteroids might have interacted with membrane molecules,reflected by the fact that certain effects can be modulatedthrough membrane receptors. In line with this suggestion isthe fact that steroids are mainly absent in bacteria orarchaea. However, the occurrence of steroids required thepresence of enzymes, capable of performing the essentialsynthetic reactions. It appears likely that these enzymeswere initially involved in xenobiotic metabolism and trans-formed into ligand-controlling `switches' with the emer-gence of receptor proteins. This is supported by the the factthat the main enzyme families involved in these trans-formation reactions, the CypP450, AKR and SDR familiesare evolutionarily old and are found in all forms of life[23,79,124,127,128]. Apparently, these enzyme familiesconstitute the oldest component of the describedbiotransformation-steroid hormone-nuclear receptor signal-ling system. However, similar functions as in highereukaryots, i.e. modulation of signalling molecules, arealso found in certain bacteria, e.g. in Rhizobia±plantinteractions requiring the SDR enzyme NodG [129].

With the occurrence of intracellular binding proteins,a transition from membrane interaction to intracellularliganding and DNA binding probably occurred. This view isconsistent with observations dating the appearence ofnuclear receptors to early metazoan life. At this time, thebasic components of the transcription machinery hadevolved, allowing to impose a regulatory system such astranscription factors. Subsequently, gene duplication, trans-fer and mutational events allowed cellular life to adapt tomany different conditions and environmental compounds[123,130].

C O N C L U S I O N S A N D P E R S P E C T I V E S

In this review we have presented examples illustratingan ancient principle of cellular communication found in

4120 S. Nobel et al. (Eur. J. Biochem. 268) q FEBS 2001

Page 9: Metabolic conversion as a pre-receptor control mechanism for lipophilic hormones

vertebrates and eukaryots, e.g. from Drosophila and C.elegans to humans, even with rudimentary roots in bacteria.The common theme for the different types of hormones andmediators, is that they are shuttled between biologicallyactive and inactive forms by enzymes that belong to a fewdifferent superfamilies. Thus, the important physiologicalprinciple is that the hormonal action in a target cell isdetermined by enzyme action within the cell, not only bycirculating hormone levels and receptor expression. Theemerging question is why such a complex system betweendifferent ligands, enzymes and receptors has evolved?Local biotransformation of the ligand allows a givenreceptor to be utilized for different ligands in differentcells (e.g. MR in CNS and kidney, binding cortisol/aldosterone). It clearly is a way of evolution to introducespecificity while maintaining flexibility for adaptation toexternal signals and stressors, i.e. by combination of twoserial semispecific building blocks (enzymes, receptors)the system achieves higher specificity within a givencompartment through a filter effect.

Modification of nuclear receptor signalling pathways is acurrently pursued avenue in drug development, exemplifiedwith the development of PPAR a and g agonists [131,132].As outlined here, it is possible to find alternative drugtargets more `upstream' in the signalling pathway. Indeed, anumber of enzymes described in this review are already infocus in drug development (e.g. 3a-HSD or 17b-HSD inhormone-dependent cancer forms [133,134]), while othersare waiting to be explored [135].

All, or most, of the molecules described in this reviewhave long been recognized as signaling molecules, althoughthe signal transduction principles have been deduced later.A current research trend indicates the reverse: a signalingsystem with the key molecular component waiting tobecome identified, which is the case for some of the`orphan nuclear receptors' [4]. This search has lead to theidentification of signaling molecules separate from classicalhormones [136±139]. Furthermore, it is now widelyaccepted that important signaling molecules are derivedfrom major food constituents, such as polyunsaturated fattyacids [140±142]. Many enzymes and isoforms involved intransformation of these and other nonclassical signalingmolecules are waiting to be identified, and it would not besurprising if the principle of enzymatic modulation of localeffects will be extended into further areas of signalingmolecules.

A C K N O W L E D G E M E N T S

Critical reading of the manuscript and fruitful discussions with Jan

SjoÈvall, Jesper HaeggstroÈm and Anders Berkenstam, Stockholm are

gratefully acknowledged. Research in the authors' laboratories has

been supported by the European Commission (BIO4CT2123), Swedish

Union of Physicians, NovoNordisk Foundation, Loo och Hans Oster-

mann Foundation, Pharmacia Corporation, BioNetWorks GmbH,

Munich, Stockholm University, and Karolinska Institutet.

R E F E R E N C E S

1. Beato, M., Herrlich, P. & SchuÈtz, G. (1995) Steroid hormone

receptors: many actors in search of a plot. Cell 83, 851±857.

2. Mangelsdorf, D., Thummel, C., Beato, M., Herrlich, P., Schutz,

G., Umesono, K., Blumberg, B., Kastner, P., Mark, M. &

Chambon, P. (1995) The nuclear receptor superfamily: the second

decade. Cell 83, 835±839.

3. Lemon, B.D. & Freedman, L.P. (1999) Nuclear receptor cofactors

as chromatin remodelers. Curr. Opin. Genet Dev. 9, 499±504.

4. Kliewer, S.A., Lehmann, J.M. & Willson, T.M. (1999) Orphan

nuclear receptors: shifting endocrinology into reverse. Science

284, 757±760.

5. Nuclear receptor nomenclature committee. (1999) A unified

nomenclature system for the nuclear receptor superfamily. Cell

97, 161±163.

6. Teoh, H. & Man, R.Y. (2000) Enhanced relaxation of porcine

coronary arteries after acute exposure to a physiological level of

17b-estradiol involves non-genomic mechanisms and the cyclic

AMP cascade. Br. J Pharmacol. 129, 1739±1747.

7. McKenna, N.J., Xu, J., Nawaz, Z., Tsai, S.Y., Tsai, M.J. &

O'Malley, B.W. (1999) Nuclear receptor coactivators: multiple

enzymes, multiple complexes, multiple functions. J. Steroid

Biochem. Mol. Biol. 69, 3±12.

8. Xu, L., Glass, C.K. & Rosenfeld, M.G. (1999) Coactivator and

corepressor complexes in nuclear receptor function. Curr. Opin.

Genet. Dev. 9, 140±147.

9. Freedman, L.P. (1999) Increasing the complexity of coactivation

in nuclear receptor signaling. Cell 97, 5±8.

10. McKenna, N.J., Lanz, R.B. & O'Malley, B.W. (1999) Nuclear

receptor coregulators: cellular and molecular biology. Endocr.

Rev. 20, 321±344.

11. Jenster, G. (1998) Coactivators and corepressors as mediators of

nuclear receptor function: an update. Mol. Cell. Endocrinol. 143,

1±7.

12. Pearce, D. (1994) A mechanistic basis for distinct mineralo-

corticoid and glucocorticoid receptor transcriptional specificities.

Steroids 59, 153±159.

13. Salbert, G., Fanjul, A., Piedrafita, F.J., Lu, X.P., Kim, S.J., Tran,

P. & Pfahl, M. (1993) Retinoic acid receptors and retinoid X

receptor-a down-regulate the transforming growth factor-b 1

promoter by antagonizing AP-1 activity. Mol. Endocrinol. 7,

1347±1356.

14. Takeda, T., Kurachi, H., Yamamoto, T., Nishiom, Y., Nakatsuji,

Y., Morishige, K., Miyake, A. & Murata, Y. (1998) Crosstalk

between the interleukin-6 (IL-6)-JAK-STAT and the

glucocorticoid-nuclear receptor pathway: synergistic activation

of IL-6 response element by IL-6 and glucocorticoid.

J. Endocrinol. 159, 323±330.

15. GoÈttlicher, M., Heck, S. & Herrlich, P. (1998) Transcriptional

cross-talk, the second mode of steroid hormone receptor action.

J. Mol. Med. 76, 480±489.

16. Caelles, C., Gonzalez-Sancho, J.M. & Munoz, A. (1997) Nuclear

hormone receptor antagonism with AP-1 by inhibition of the JNK

pathway. Genes Dev. 11, 3351±3364.

17. Helmberg, A., Auphan, N., Caelles, C. & Karin, M. (1995)

Glucocorticoid-induced apoptosis of human leukemic cells is

caused by the repressive function of the glucocorticoid receptor.

EMBO J. 14, 452±460.

18. Bodwell, J.E., Webster, J.C., Jewell, C.M., Cidlowski, J.A., Hu,

J.M. & Munck, A. (1998) Glucocorticoid receptor phosphory-

lation: overview, function and cell cycle-dependence. J. Steroid

Biochem. Mol Biol. 65, 91±99.

19. Smith, C.L. (1998) Cross-talk between peptide growth factor and

estrogen receptor signaling pathways. Biol. Reprod. 58, 627±632.

20. Labrie, F. (1991) Intracrinology. Mol. Cell. Endocrinol. 78,

C113±C118.

21. Labrie, F., BeÂlanger, A., Simard, J., Luu-The, V. & Labrie, C.

(1995) DHEA and peripheral androgen and estrogen formation:

intracrinology. Ann. NY Acad. Sci. 774, 16±28.

22. Labrie, F., Luu-The, V., Lin, S.X., Simard, J., Labrie, C., El-Alfy,

M., Pelletier, G. & Belanger, A. (2000) Intracrinology: role of the

q FEBS 2001 Enzymatic control of lipid hormones (Eur. J. Biochem. 268) 4121

Page 10: Metabolic conversion as a pre-receptor control mechanism for lipophilic hormones

family of 17b-hydroxysteroid dehydrogenases in human physiol-

ogy and disease. J. Mol. Endocrinol. 25, 1±16.

23. JoÈrnvall, H., Persson, B., Krook, M., Atrian, S., Gonzalez-Duarte,

R., Jeffery, J. & Ghosh, D. (1995) Short-chain dehydrogenases/

reductases (SDR). Biochemistry 34, 6003±6013.

24. Oppermann, U.C.T., Persson, B., Filling, C. & JoÈrnvall, H. (1997)

Structure function relationships of hydroxysteroid dehydro-

genases. Adv. Exp. Med. Biol. 414, 403±415.

25. Penning, T.M. (1997) Molecular endocrinology of hydroxy-

steroid dehydrogenases. Endocr. Rev. 18, 281±305.

26. New, M.I. & White, P.C. (1995) Genetic disorders of steroid

hormone synthesis and metabolism. Baillieres Clin. Endocrinol.

Metab. 9, 525±554.

27. Baulieu, E.E. & Robel, P. (1995) Non-genomic mechanisms of

action of steroid hormones. Ciba Found Symp 191, 24±42.

28. Pazirandeh, A., Xue, Y., Rafter, I., SjoÈvall, J., Jondal, M. & Okret,

S. (1999) Paracrine glucocorticoid activity produced by mouse

thymic epithelial cells. FASEB J. 13, 893±901.

29. Silvestre, J.S., Robert, V., Heymes, C., Aupetit-Faisant, B.,

Mouas, C., Moalic, J.M., Swynghedauw, B. & Delcayre, C.

(1998) Myocardial production of aldosterone and corticosterone

in the rat. Physiological regulation. J. Biol. Chem. 273,

4883±4891.

30. Konig, A., Happle, R., Bornholdt, D., Engel, H. & Grzeschik,

K.H. (2000) Mutations in the NSDHL gene, encoding a

3b-hydroxysteroid dehydrogenase, cause CHILD syndrome.

Am. J. Med. Genet. 90, 339±346.

31. Geissler, W.M., Davis, D.L., Wu, L., Bradshaw, K.D., Patel, S.,

Mendonca, B.B., Elliston, K.O., Wilson, J.D., Russell, D.W. &

Andersson, S. (1994) Male pseudohermaphroditism caused by

mutations of testicular 17 b-hydroxysteroid dehydrogenase 3.

Nat. Genet. 7, 34±39.

32. Munck, A. & Naray-Fejes-Toth, A. (1992) The ups and downs of

glucocorticoid physiology. Permissive and suppressive effects

revisited. Mol. Cell. Endocrinol. 90, C1±C4.

33. Burton, P.J. & Waddell, B.J. (1999) Dual function of

11b-hydroxysteroid dehydrogenase in placenta: modulating

placental glucocorticoid passage and local steroid action. Biol.

Reprod. 60, 234±240.

34. Brooks, R.V. (1979) Biosynthesis and metabolism of adreno-

corticol steroids. In The Adrenal Gland (James, V. H. T., ed.), pp.

67±92. Raven Press, New York.

35. Marver, D. (1984) Assessment of mineralocorticoid activity in

the rabbit colon. Am. J. Physiol. 246, F437±F446.

36. Hollenberg, S.M., Weinberger, C., Ong, E.S., Cerelli, G., Oro, A.,

Lebo, R., Thompson, E.B., Rosenfeld, M.G. & Evans, R.M.

(1985) Primary structure and expression of a functional human

glucocorticoid receptor cDNA. Nature 318, 635±641.

37. Arriza, J.L., Weinberger, C., Cerelli, G., Glaser, T.M., Handelin,

B.L., Housman, D.E. & Evans, R.M. (1987) Cloning of human

mineralocorticoid receptor complementary DNA: structural and

functional kinship with the glucocorticoid receptor. Science 237,

268±275.

38. Ulick, S., Levine, L.S., Gunczler, P., Zanconato, G., Ramirez,

L.C., Rauh, W., Rosler, A., Bradlow, H.L. & New, M.I. (1979) A

syndrome of apparent mineralocorticoid excess associated with

defects in the peripheral metabolism of cortisol. J. Clin.

Endocrinol Metab. 49, 757±764.

39. Mune, T., Rogerson, F.M., Nikkila, H., Agarwal, A.K. & White,

P.C. (1995) Human hypertension caused by mutations in the

kidney isozyme of 11 b-hydroxysteroid dehydrogenase. Nat

Genet. 10, 394±399.

40. Kotelevtsev, Y., Brown, R.W., Fleming, S., Kenyon, C., Edwards,

C.R.W., Seckl, J.R. & Mullins, J.J. (1999) Hypertension in mice

lacking 11b-hydroxysteroid dehydrogenase type 2. J. Clin.

Invest. 103, 683±689.

41. Funder, J.W. (2000) Aldosterone and mineralocorticoid receptors:

orphan questions. Kidney Int. 57, 1358±1363.

42. Spencer, R.L., Kim, P.J., Kalman, B.A. & Cole, M.A. (1998)

Evidence for mineralocorticoid receptor facilitation of gluco-

corticoid receptor-dependent regulation of hypothalamic-

pituitary-adrenal axis activity. Endocrinology 139, 2718±2726.

43. Stewart, P.M. & Krozowski, Z.S. (1999) 11b-Hydroxysteroid

dehydrogenase. Vitam. Horm. 57, 249±324.

44. Jamieson, P.M., Chapman, K.E., Edwards, C.R. & Seckl, J.R.

(1995) 11 b-hydroxysteroid dehydrogenase is an exclusive 11 b-

reductase in primary cultures of rat hepatocytes: effect of

physicochemical and hormonal manipulations. Endocrinology

136, 4754±4761.

45. Whorwood, C.B., Franklyn, J.A., Sheppard, M.C. & Stewart,

P.M. (1992) Tissue localization of 11b-hydroxysteroid dehydro-

genase and its relationship to the glucocorticoid receptor.

J. Steroid Biochem. Mol. Biol. 41, 21±28.

46. Kotelevtsev, Y., Holmes, M.C., Burchell, A., Houston, P.M.,

Schmoll, D., Jamieson, P., Best, R., Brown, R., Edwards, C.R.,

Seckl, J.R. & Mullins, J.J. (1997) 11b-hydroxysteroid dehydro-

genase type 1 knockout mice show attenuated glucocorticoid-

inducible responses and resist hyperglycemia on obesity or stress.

Proc. Natl Acad. Sci. USA 94, 14924±14929.

47. Ui, T., Mitsunaga, M., Tanaka, T. & Horiguchi, M. (1982)

Determination of prednisone and prednisolone in human serum

by high- performance liquid chromatography ± especially on

impaired conversion of corticosteroids in patients with chronic

liver disease. J. Chromatogr. 239, 711±716.

48. Madsbad, S., Bjerregaard, B., Henriksen, J.H., Juhl, E. & Kehlet,

H. (1980) Impaired conversion of prednisone to prednisolone in

patients with liver cirrhosis. Gut 21, 52±56.

49. NordenstroÈm, A., Marcus, C., Axelson, M., Wedell, A. & RitzeÂn,

E.M. (1999) Clinical case seminar. Failure of cortisone acetate

treatment in congenital adrenal hyperplasia because of defective

11b-hydroxysteroid dehydrogenase reductase activity. J. Clin.

Endocrinol Metab. 84, 1210±1213.

50. Jamieson, A., Wallace, A.M., Andrew, R., Nunez, B.S., Walker,

B.R., Fraser, R., White, P.C. & Connell, J.M.C. (1999) Apparent

cortisone reductase deficiency: a functional defect in 11b-

hydroxysteroid dehydrogenase type 1. J. Clin. Endocrinol

Metab. 84, 3570±3574.

51. Andrews, R.C. & Walker, B.R. (1999) Glucocorticoids and

insulin resistance: old hormones, new targets. Clin. Sci. 96,

513±523.

52. Davani, B., Khan, A., Hult, M., MaÊrtensson, E., Okret, S.,

Efendic, S., JoÈrnvall, H. & Oppermann, U.C. (2000) Type 1 11b-

hydroxysteroid dehydrogenase mediates glucocorticoid activation

and insulin release in pancreatic islets. J. Biol. Chem. 275,

34841±34844.

53. Hinson, J.P. & Raven, P.W. (1999) DHEA deficiency syndrome: a

new term for old age? J. Endocrinol. 163, 1±5.

54. Strott, C.A. (1996) Steroid sulfotransferases. Endocrine Rev. 17,

670±697.

55. Wang, D.Y. & Bulbrook, R.D. (1967) The metabolic clearance

rates of dehydroepiandrosterone, testosterone and their sulfate

esters in man, rat and rabbit. J. Endocrinol. 38, 307±318.

56. Nagata, K. & Yamazoe, Y. (2000) Pharmacogenetics of

sulfotransferase. Annu. Rev. Pharmacol Toxicol. 40, 159±176.

57. Reed, M.J., Purohit, A., Duncan, L.J., Singh, A., Roberts, C.J.,

Williams, G.J. & Potter, B.V.L. (1995) The role of cytokines and

sulfatase inhibitors in regulating oestrogen synthesis in breast

tumours. J. Steroid Biochem. Molec. Biol. 53, 413±420.

58. Parenti, G., Meroni, G. & Ballabio, A. (1997) The sulfatase gene

family. Curr. Opin. Genet Dev. 7, 386±391.

59. Henry, F.J. & Bassett, J.R. (1985) Corticosterone storage

within the adrenal cortex: evidence for a sulfate conjugate.

J. Endocrinol. 104, 381±386.

4122 S. Nobel et al. (Eur. J. Biochem. 268) q FEBS 2001

Page 11: Metabolic conversion as a pre-receptor control mechanism for lipophilic hormones

60. Peltoketo, H., Luu-The, V., Simard, J. & Adamski, J. (1999)

17b-hydroxysteroid dehydrogenase (HSD) /17-ketosteroid reduc-

tase (KSR) family; nomenclature and main characteristics of the

17HSD/KSR enzymes. J. Mol. Endocrinol. 23, 1±11.

61. Suzuki, T., Sasano, H., Andersson, S. & Mason, J.I. (2000)

3b-hydroxysteroid dehydrogenase/delta5!4-isomerase activity

associated with the human 17b-hydroxysteroid dehydrogenase

type 2 isoform. J. Clin. Endocrinol Metab. 85, 3669±3672.

62. Baker, M.E. (1998) Evolution of mammalian 11b- and 17b-

hydroxysteroid dehydrogenases-type 2 and retinol dehydro-

genases from ancestors in Caenorhabitis elegans and evidence

for horizontal transfer of a eukaryote dehydrogenase to E. coli.

J. Steroid Biochem. Molec. Biol. 66, 355±363.

63. Nokelainen, P., Peltoketo, H., Mustonen, M. & Vihko, P. (2000)

Expression of mouse 17b-hydroxysterid dehydrogenase/17-

ketosterid reductase type 7 in the ovary, uterus, and placenta:

localization from implantation to late pregnancy. Endocrinology

141, 772±778.

64. English, M.A., Kane, K.F., Cruickshank, N., Langman, M.J.S.,

Stewart, P.M. & Hewison, M. (1999) Loss of estrogen

inactivation in colonic cancer. J. Clin. Endocriol. Metab. 84,

2080±2085.

65. Parish, E.J., Li, S. & Rao, Z. (2000) Design and synthesis of new

steroidal inhibitors of estrogen synthase (aromatase). Lipids 35,

271±277.

66. Kester, M.H.A., Bulduk, S., Tibboel, D., Meinl, W., Glatt, H.,

Falany, C.N., Coughtrie, M.W.H., Bergman, AÊ ., Safe, S.H.,

Kuiper, G.G.J.M., Schuur, A.G., Brouwer, A. & Visser, T.J.

(2000) Potent inhibition of estrogen sulfotrasferase by hydroxy-

lated PCB metabolites: a novel pathway explaining the estrogenic

activity of PCBs. Endocrinology 141, 1897±1900.

67. Andersson, S.H. & SjoÈvall, J. (1986) Effects of ethanol on steroid

profiles in the rat testis. Biochim. Biophys. Acta. 876, 352±357.

68. Andersson, S.H., Cronholm, T. & SjoÈvall, J. (1986) Effects of

ethanol on the levels of unconjugated and conjugated androgens

and estrogens in plasma of men. J. Steroid Biochem. 24,

1193±1198.

69. Andersson, S., Cronholm, T. & SjoÈvall, J. (1986) Redox effects

of ethanol on steroid metabolism. Alcohol Clin. Exp Res. 10,

55S±63S.

70. Andersson, S., Berman, D.M., Jenkins, E.P. & Russell, D.W.

(1991) Deletion of steroid 5 a-reductase 2 gene in male

pseudohermaphroditism. Nature 354, 159±161.

71. Penning, T.M., Burczynski, M.E., Jez, J.M., Hung, C.F., Lin,

H.K., Ma, H., Moore, M., Palackal, N. & Ratnam, K. (2000)

Human 3a-hydroxysteroid dehydrogenase isoforms (AKR1C1±

AKR1C4) of the aldo-keto reductase superfamily: functional

plasticity and tissue distribution reveals roles in the inactivation

and formation of male and female sex hormones. Biochem. J.

351, 67±77.

72. Isaacs, J.T. (1983) Changes in dihydrotestosterone metabolism

and the development of benign prostatic hyperplasia in the aging

beagle. J. Steroid Biochem. 18, 749±757.

73. He, X.Y., Merz, G., Yang, Y.Z., Pullakart, R., Mehta, P., Schulz,

H. & Yang, S.Y. (2000) Function of human brain short chain

L-3-hydroxyacyl coenzyme A dehydrogenase in androgen

metabolism. Biochim. Biophys. Acta. 1484, 267±277.

74. Axelson, M., Schumacher, G., SjoÈvall, J., Gustafsson, B. &

Lindell, J.O. (1975) Identification and quantitative determination

of steroids in bovine corpus luteum during oestrous cycle and

pregnancy. Acta Endocrinol. (Copenh). 80, 149±164.

75. SjoÈvall, K. (1970) Sulfates of pregnanolones, pregnanediols and

16-hydroxysteroids in plasma from pregnant women. Ann. Clin.

Res. 2, 409±413.

76. Zhang, Y., Dufort, I., Rheault, P. & Luu-The, V. (2000)

Characterization of a human 20a-hydroxysteroid dehydrogenase.

J. Mol. Endocrinol. 25, 221±228.

77. Albarracin, C.T., Parmer, T.G., Duan, W.R., Nelson, S.E. &

Gibori, G. (1994) Identification of a major prolactin-regulated

protein as 20 a-hydroxysteroid dehydrogenase: coordinate regu-

lation of its activity, protein content, and messenger ribonucleic

acid expression. Endocrinology 134, 2453±2460.

78. Quinkler, M., Johanssen, S., Bumke-Vogt, C., Oelkers, W., Bahr,

V. & Diederich, S. (2001) Enzyme-mediated protection of the

mineralocorticoid receptor against progesterone in the human

kidney. Mol. Cell Endocrinol. 171, 21±24.

79. Jez, J.M., Flynn, T.G. & Penning, T.M. (1997) A new nomen-

clature for the aldo-keto reductase superfamily. Biochem.

Pharmacol. 54, 639±647.

80. Zhong, L., Ou, J., Barkai, U., Mao, J.F., Frasor, J. & Gibori, G.

(1998) Molecular cloning and characterization of the rat ovarian

20 a-hydroxysteroid dehydrogenase gene. Biochem. Biophys.

Res. Commun. 249, 797±803.

81. Mensah-Nyagan, A.G., Do-Rego, J.L., Beaujean, D., Luu-The,

V., Pelletier, G. & Vaudry, H. (1999) Neurosteroids: expression of

steroidogenic enzymes and regulation of steroid biosynthesis in

the central nervous system. Pharmacol. Rev. 51, 63±81.

82. Corpechot, C., Robel, P., Axelson, M., Sjovall, J. & Baulieu, E.E.

(1981) Characterization and measurement of dehydroepiandro-

sterone sulfate in rat brain. Proc. Natl Acad. Sci. USA 78,

4704±4707.

83. Majewska, M.D., Harrison, N.L., Schwartz, R.D., Barker, J.L. &

Paul, S.M. (1986) Steroid hormone metabolites are barbiturate-

like modulators of the GABAA receptor. Science. 232, 1004±1007.

84. Smith, S.S., Gong, Q.H., Hsu, F.-C., Markowitz, R.S., French-

Mullen, J.M.H. & Li, X. (1998) GABAA receptor a4 subunit

suppression prevents withdrawal properties of an endogenous

steroid. Nature 392, 926±930.

85. Paul, S.M. & Purdy, R.H. (1992) Neuroactive steroids. FASEB J.

6, 2311±2322.

86. Grazzini, E., Guillon, G., Mouillac, B. & Zingg, H.H. (1998)

Inhibition of oxytocin receptor function by direct binding of

progesterone. Nature 392, 509±512.

87. Baulieu, E.E. (2000) New' active steroids and an unforeseen

mechanism of action. CR Acad. Sci. III (323), 513±518.

88. Brown, A.J., Dusso, A. & Slatopolsky, E. (1999) Vitamin D. Am. J

Physiol. 277, F157±F175.

89. Wikvall, K. (2001) Cytochrome P450 enzymes in the bio-

activation of vitamin D to its hormonal form. Int. J. Mol. Med. 7,

201±209.

90. Penna, G. & Adorini, L. (2000) 1-Alpha,25-dihydroxyvitamin D3

inhibits differentiation, maturation, activation, and survival of

dendritic cells leading to impaired alloreactive T cell activation.

J. Immunol. 164, 2405±2411.

91. Hager, G., Formanek, M., Gedlicka, C., Thurnher, D., Knerer, B.

& Kornfehl, J. (2001) 1,25(OH)2 vitamin D3 induces elevated

expression of the cell cycle-regulating genes P21 and P27 in

squamous carcinoma cell lines of the head and neck. Acta

Otolaryngol. 121, 103±109.

92. Brown, M.A., Haughton, M.A., Grant, S.F., Gunnell, A.S.,

Henderson, N.K. & Eisman, J.A. (2001) Genetic control of bone

density and turnover: role of the collagen 1a1, estrogen receptor,

and vitamin D receptor genes. J. Bone Miner. Res. 16, 758±764.

93. Rook, G.A., Hernandez-Pando, R. & Lightman, S.L. (1994)

Hormones, peripherally activated prohormones and regulation of

the Th1/ Th2 balance. Immunol. Today 15, 301±303.

94. Wilckens, T. & De Rijk, R. (1997) Glucocorticoids and immune

function: unknown dimensions and new frontiers. Immunol.

Today 18, 418±424.

95. Wilckens, T. (1995) Glucocorticoids and immune function:

physiological relevance and pathogenic potential of hormonal

dysfunction. Trends Pharmacol. Sci. 16, 193±197.

96. Chambon, P. (1996) A decade of molecular biology of retinoic

acid receptors. FASEB J. 10, 940±954.

q FEBS 2001 Enzymatic control of lipid hormones (Eur. J. Biochem. 268) 4123

Page 12: Metabolic conversion as a pre-receptor control mechanism for lipophilic hormones

97. Noy, N. (2000) Retinoid-binding proteins: mediators of retinoid

action. Biochem. J. 348, 481±495.

98. Napoli, J.L. (1999) Interactions of retinoid binding proteins and

enzymes in retinoid metabolism. Biochim. Biophys. Acta 1440,

139±162.

99. Goodman, A.B. (1998) Three independent lines of evidence

suggest retinoids as causal to schizophrenia. Proc. Natl Acad. Sci.

USA 95, 7240±7244.

100. Napoli, J.L. (2000) A gene knockout corroborates the integral

function of cellular retinol-binding protein in retinoid metabol-

ism. Nutr. Rev. 58, 230±236.

101. Crosas, B., Allali-Hassani, A., Martinez, S.E., Martras, S.,

Persson, B., JoÈrnvall, H., Pares, X. & Farres, J. (2000) Molecular

basis for differential substrate specificity in class IV alcohol

dehydrogenases: a conserved function in retinoid metabolism but

not in ethanol oxidation. J. Biol. Chem. 275, 25180±25187.

102. Duester, G. (2000) Families of retinoid dehydrogenases regulat-

ing vitamin A function: production of visual pigment and retinoic

acid. Eur. J. Biochem. 267, 4315±4324.

103. Ghyselinck, N.B., Bavik, C., Sapin, V., Mark, M., Bonnier, D.,

Hindelang, C., Dierich, A., Nilsson, C.B., Hakansson, H.,

Sauvant, P., Azais-Braesco, V., Frasson, M., Picaud, S. &

Chambon, P. (1999) Cellular retinol-binding protein I is essential

for vitamin A homeostasis. EMBO J. 18, 4903±4914.

104. Deltour, L., Foglio, M.H. & Duester, G. (1999) Impaired retinol

utilization in Adh4 alcohol dehydrogenase mutant mice. Dev

Genet. 25, 1±10.

105. Deltour, L., Foglio, M.H. & Duester, G. (1999) Metabolic

deficiencies in alcohol dehydrogenase Adh1, Adh3, and Adh4

null mutant mice. Overlapping roles of Adh1 and Adh4 in ethanol

clearance and metabolism of retinol to retinoic acid. J. Biol.

Chem. 274, 16796±16801.

106. Chai, X., Boerman, M.H., Zhai, Y. & Napoli, J.L. (1995) Cloning

of a cDNA for liver microsomal retinol dehydrogenase. A tissue-

specific, short-chain alcohol dehydrogenase. J. Biol. Chem. 270,

3900±3904.

107. Biswas, M.G. & Russell, D.W. (1997) Expression cloning and

characterization of oxidative 17b- and 3a-hydroxysteroid

dehydrogenases from rat and human prostate. J. Biol. Chem.

272, 15959±15966.

108. Gough, W.H., VanOoteghem, S., Sint, T. & Kedishvili, N.Y.

(1998) cDNA cloning and characterization of a new human

microsomal NAD1-dependent dehydrogenase that oxidizes

all-trans-retinol and 3a-hydroxysteroids. J. Biol. Chem. 273,

19778±19785.

109. Su, J., Lin, M. & Napoli, J.L. (1999) Complementary

deoxyribonucleic acid cloning and enzymatic characterization

of a novel 17b/3a-hydroxysteroid/retinoid short chain dehydro-

genase/reductase. Endocrinology 140, 5275±5284.

110. Perozich, J., Nicholas, H., Wang, B.C., Lindahl, R. & Hempel, J.

(1999) Relationships within the aldehyde dehydrogenase

extended family. Protein Sci. 8, 137±146.

111. Kelly, G.S. (2000) Peripheral metabolism of thyroid hormones:

a review. Altern. Med. Rev. 5, 306±333.

112. Asteria, C. (1998) Crucial role for type II iodothyronine

deiodinase in the metabolic coupling between glial cells and

neurons during brain development. Eur. J. Endocrinol. 138,

370±371.

113. Croteau, W., Davey, J.C., Galton, V.A. & St Germain, D.L.

(1996) Cloning of the mammalian type II iodothyronine

deiodinase. A selenoprotein differentially expressed and regu-

lated in human and rat brain and other tissues. J. Clin. Invest. 98,

405±417.

114. Mandel, S.J., Berry, M.J., Kieffer, J.D., Harney, J.W., Warne, R.L.

& Larsen, P.R. (1992) Cloning and in vitro expression of the

human selenoprotein, type I iodothyronine deiodinase. J. Clin.

Endocrinol. Metab. 75, 1133±1139.

115. Salvatore, D., Low, S.C., Berry, M., Maia, A.L., Harney, J.W.,

Croteau, W., St. Germain, D.L. & Larsen, P.R. (1995) Type 3

lodothyronine deiodinase: cloning, in vitro expression, and

functional analysis of the placental selenoenzyme. J. Clin. Invest.

96, 2421±2430.

116. Bates, J.M., St. Germain, D.L. & Galton, V.A. (1999) Expression

profiles of the three iodothyronine deiodinases, D1, D2, and D3,

in the developing rat. Endocrinology 140, 844±851.

117. St Germain, D.L. (1999) Development effects of thyroid

hormone: the role of deiodinases in regulatory control. Biochem.

Soc. Trans. 27, 83±88.

118. Serhan, C.N., Haeggstrom, J.Z. & Leslie, C.C. (1996) Lipid

mediator networks in cell signaling: update and impact of

cytokines. FASEB J. 10, 1147±1158.

119. HaeggstroÈm, J.Z. (2000) Structure, function, and regulation of

leukotriene A4 hydrolase. Am. J. Respir. Crit. Care Med. 161,

S25±S31.

120. Krook, M., Marekov, L. & JoÈrnvall, H. (1990) Purification

and structural characterization of placental NAD(1)-linked

15-hydroxyprostaglandin dehydrogenase. The primary structure

reveals the enzyme to belong to the short-chain alcohol

dehydrogenase family. Biochemistry 29, 738±743.

121. Ensor, C.M., Yang, J.Y., Okita, R.T. & Tai, H.H. (1990) Cloning

and sequence analysis of the cDNA for human placental

NAD(1)-dependent 15-hydroxyprostaglandin dehydrogenase.

J. Biol. Chem. 265, 14888±14891.

122. Pugliese, G., Spokas, E.G., Marcinkiewicz, E. & Wong, P.Y.

(1985) Hepatic transformation of prostaglandin D2 to a new

prostanoid, 9a,11b-prostaglandin F2, that inhibits platelet

aggregation and constricts blood vessels. J. Biol. Chem. 260,

14621±14625.

123. Owen, G.I. & Zelent, A. (2000) Origins and evolutionary

diversification of the nuclear receptor superfamily. Cell Mol.

Life Sci. 57, 809±827.

124. Gonzalez, F.J. & Nebert, D.W. (1990) Evolution of the P450 gene

superfamily: animal-plant `warfare', molecular drive and human

genetic differences in drug oxidation. Trends Genet. 6, 182±186.

125. Karlson, P. (1983) Eighth Adolf Butenandt lecture. Why are so

many hormones steroids? Hoppe Seylers Z. Physiol. Chem. 364,

1067±1087.

126. Incardona, J.P. & Eaton, S. (2000) Cholesterol in signal

transduction. Curr. Opin. Cell Biol. 12, 193±203.

127. Jez, J.M., Bennett, M.J., Schlegel, B.P., Lewis, M. & Penning,

T.M. (1997) Comparative anatomy of the aldo-keto reductase

superfamily. Biochem. J. 326, 625±636.

128. Baker, M.E. (1990) A common ancestor for human placental 17

b-hydroxysteroid dehydrogenase, Streptomyces coelicolor actIII

protein, and Drosophila melanogaster alcohol dehydrogenase.

FASEB J. 4, 222±226.

129. Baker, M.E. (1992) Evolution of regulation of steroid-mediated

intercellular communication in vertebrates: insights from flavo-

noids, signals that mediate plant-rhizobia symbiosis. J. Steroid

Biochem. Mol. Biol. 41, 301±308.

130. Baker, M.E. (1997) Steroid receptor phylogeny and vertebrate

origins. Mol. Cell Endocrinol. 135, 101±107.

131. Kersten, S., Desvergne, B. & Wahli, W. (2000) Roles of PPARs in

health and disease. Nature 405, 421±424.

132. Murphy, G.J. & Holder, J.C. (2000) PPAR-gamma agonists:

therapeutic role in diabetes, inflammation and cancer. Trends

Pharmacol. Sci. 21, 469±474.

133. Pasqualini, J.R., Ebert, C. & Chetrite, G.S. (1999) The SEEM:

selective estrogen enzyme modulators in breast cancer. Gynecol.

Endocrinol. 13(Suppl. 6), 1±8.

134. Henderson, B.E. & Feigelson, H.S. (2000) Hormonal carcino-

genesis. Carcinogenesis 21, 427±433.

4124 S. Nobel et al. (Eur. J. Biochem. 268) q FEBS 2001

Page 13: Metabolic conversion as a pre-receptor control mechanism for lipophilic hormones

135. Duax, W.L., Ghosh, D. & Pletnev, V. (2000) Steroid dehydro-

genase structures, mechanism of action, and disease. Vitam.

Horm. 58, 121±148.

136. Wei, P., Zhang, J., Egan-Hafley, M., Liang, S. & Moore, D.D.

(2000) The nuclear receptor CAR mediates specific xenobiotic

induction of drug metabolism. Nature 407, 920±923.

137. Chawla, A., Saez, E. & Evans, R.M. (2000) Don't know much

bile-ology. Cell 103, 1±4.

138. Russell, D.W. (2000) Oxysterol biosynthetic enzymes. Biochim.

Biophys. Acta. 1529, 126±135.

139. Peet, D.J., Janowski, B.A. & Mangelsdorf, D.J. (1998) The

LXRs: a new class of oxysterol receptors. Curr. Opin. Genet.

Dev. 8, 571±575.

140. Jump, D.B., Thelen, A., Ren, B. & Mater, M. (1999) Multiple

mechanisms for polyunsaturated fatty acid regulation of hepatic

gene transcription. Prostaglandins Leukot. Essent. Fatty Acids 60,

345±349.

141. de Urquiza, A.M., Liu, S., SjoÈberg, M., Zetterstrom, R.H.,

Griffiths, W., SjoÈvall, J. & Perlmann, T. (2000) Docosahexaenoic

acid, a ligand for the retinoid X receptor in mouse brain. Science

290, 2140±2144.

142. Banner, C.D., Gottlicher, M., Widmark, E., SjoÈvall, J., Rafter, J.J.

& Gustafsson, J.A. (1993) A systematic analytical chemistry/cell

assay approach to isolate activators of orphan nuclear receptors

from biological extracts: characterization of peroxisome

proliferator-activated receptor activators in plasma. J. Lipid.

Res. 34, 1583±1591.

q FEBS 2001 Enzymatic control of lipid hormones (Eur. J. Biochem. 268) 4125