You are on page 1of 9

Available online at www.sciencedirect.

com

Journal of Steroid Biochemistry & Molecular Biology 110 (2008) 39–47

Structure–activity relationships of synthetic progestins in


a yeast-based in vitro androgen bioassay
L. McRobb a,b , D.J. Handelsman c , R. Kazlauskas d , S. Wilkinson e ,
M.D. McLeod e , A.K. Heather a,b,∗
a Heart Research Institute, Sydney, NSW 2050, Australia
b Discipline of Medicine, University of Sydney, NSW 2006, Australia
c ANZAC Research Institute, University of Sydney, Sydney, NSW 2139, Australia
d National Measurement Institute, Pymble, NSW 2073, Australia
e School of Chemistry, F11, The University of Sydney, NSW 2006, Australia

Received 4 June 2007; accepted 2 October 2007

Abstract
The recent identification of tetrahydrogestrinone (THG), a non-marketed designer androgen used for sports doping but previously undetectable
by established mass spectrometry-based urine drug screens, and its production by a facile chemical modification of gestrinone has raised concerns
about the risks of developing designer androgens from numerous marketed progestins. We therefore have used yeast-based in vitro androgen and
progesterone bioassays to conduct a structure–activity study assessing the intrinsic androgenic potential of commercially available progestins and
their derivatives, to identify those compounds or structures with the highest risk of forming a basis for such misapplication. Progestins had a wide
range of androgenic bioactivity that was not reliably predicted for individual steroids by their progestin bioactivity. 17␣-Hydroxyprogesterone
and 19-norprogesterone derivatives with their bulky 17␤-substituents were strong progestins but generally weak androgens. 17␣-Ethynylated
derivatives of testosterone, 19-nortestosterone and 18-methyl-19-nortestosterone such as gestrinone, ethisterone, norethisterone and norgestrel had
the most significant intrinsic androgenicity of all the commercially marketed progestins. Facile chemical modification of the 17␣-ethynyl group of
each of these progestins produces 17␣-methyl, ethyl and allyl derivatives, including THG and norbolethone, which further enhanced androgenic
bioactivity. Thus by using the rapid and sensitive yeast bioassay we have screened a comprehensive set of progestins and associated structures and
identified the ethynylated testosterone, 19-nortestosterone and 18-methyl-19-nortestosterone derivatives as possessing the highest risk for abuse
and potential for conversion to still more potent androgens. By contrast, modern progestins such as progesterone, 17␣-hydroxyprogesterone and
19-norprogesterone derivatives had minimal androgenic bioactivity and pose low risk.
© 2008 Elsevier Ltd. All rights reserved.

Keywords: Progestin; Androgen receptor; Progesterone receptor; Steroid; Hormone; Androgen; Sports doping

1. Introduction tom manufactured, illicit designer androgens. The banned list of


synthetic androgens however specifies a finite list of known syn-
The recent identification of previously unknown (tetrahydro- thetic androgens whereas structural modifications using known
gestrinone or THG) [1–4] or non-marketed (norbolethone and chemistry could produce a wide array of androgenic agonists.
madol) [5,6] but potent androgens intended for use in sports For example, the identification of THG [2–4], a minor one-step
doping but previously not detectable by targeted gas chromatog- chemical modification of the commercially marketed progestin,
raphy mass spectrometry-based urine drug screens has raised gestrinone, created a potent androgen highlighting that cus-
concern regarding the potential for further development of cus- tom modification of commercially available steroids employing
known chemical transformations may produce novel potent
androgens as sports doping agents that are undetectable by estab-
∗ Corresponding author at: Heart Research Institute, 114 Pyrmont Bridge
lished screening tests. One of the most likely sources of starting
Road, Camperdown, NSW 2050, Sydney, Australia. Tel.: +61 2 8208 8900;
material are the synthetic progestins, many marketed for female
fax: +61 2 9565 5584. contraceptive, hormone replacement therapy or other gyneco-
E-mail address: heathera@hri.org.au (A.K. Heather). logical treatments. Although MS-based detection methods can

0960-0760/$ – see front matter © 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jsbmb.2007.10.008
40 L. McRobb et al. / Journal of Steroid Biochemistry & Molecular Biology 110 (2008) 39–47

be rapidly and readily developed once novel designer androgens point for further modification and development of illicit sports
are structurally identified, it would be helpful to identify which doping agents.
progestins with highest intrinsic androgenic bioactivity are of
most concern for such conversion to potent androgens. 2. Materials and methods
Currently marketed progestins fall into two main classes
according to their structural relatedness to progesterone or 2.1. Materials
19-nortestosterone [7]. Historically, this split arose from the
finding that, whereas 19-nortestosterone (nandrolone) was a All steroids used in this study were obtained from Steraloids,
more potent androgen than testosterone, it was orally inac- Sigma or were custom synthesized. 13␤-Ethyl-17␤-hydro-
tive and the introduction of a 17␣-ethynyl group to render xygon-4-en-3-one (18-methylnandrolone) and 13␤-ethyl-
19-nortestosterone orally active, as it does for estradiol, pro- 17␤-hydroxy-18,19-dinorpregn-4-en-3-one (norbolethone)
duced unexpected but highly potent progestin activity [8]. were synthesized by Dr. R. Davey and 17␤-hydroxy
Subsequently 19-norprogesterone derivatives were developed -17␣-(2-propenyl)-estr-4-en-3-one (17␣-allylnandrolone),
to reduce androgenic activity (as a female contraceptive side- 13␤-ethyl-17␤-hydroxy-17␣-(2-propenyl)-gon-4-en-3-one
effect) while enhancing progestin potency [9]. (17␣-allyl-18-methylnandrolone) and 13␤-ethyl-17␤-hydroxy-
Both androgen and progestin effects are mediated by 17␣-methylgon-4-en-3-one (17␣,18-dimethylnandrolone) by
interaction with their cognate receptors, members of the ligand- Mr. S. Wilkinson at the School of Chemistry, University of
activated nuclear receptor superfamily of genes and proteins Sydney, Australia. Steroids were dissolved in 100% (v/v)
[10]. Among the classical steroid receptor members of this ethanol.
class, the androgen and progesterone receptors have the highest
homology at both the gene and protein level, and the great- 2.2. Plasmids and reporter gene constructs
est structural homology between their respective ligand-binding
domains (LBDs) which dictate ligand specificity [11,12]. The The full-length hPRA-cDNA expression plasmid that
effects of synthetic substituents in steroids frequently broaden encodes the human PR receptor fused to the CUP1 metalloth-
specificity but are not strictly predictable. Thus although 19- ionein promoter and the PRE-B-galactosidase reporter plasmid
nortestosterone derivatives would be expected to have some that has steroid responsive elements upstream of lacZ, the ␤-
androgenic activity, this varies between synthetic steroids, and galactosidase reporter gene, were kindly provided by Dr. D.P.
conversely even those derived from 17-hydroxyprogesterone McDonnell. Yeast strain YPH500 (Mat␣, ura3-52, lys2-801,
may have androgenic activity. To date, gestrinone is the only ade2-101, trp1-63, his3-200, leu2-1) was co-transformed
marketed progestin on the World Anti-Doping Agency (WADA) with both plasmids by standard alkali-transformation (Alkali
prohibited list [13]. Yet among commercially marketed pro- Cation yeast transformation kit, BIO101 systems, Qbiogene
gestins, others have similar or greater androgenic potential than Inc., Carlsbad, CA, USA). Co-transformed yeast cells were
gestrinone [14] and therefore could serve as starting materi- selected by tryptophan and uracil auxotrophy. Yeast strain
als for conversion into potent, undetectable designer androgens. YPH500 transformed with YEPAR and YppG2 was also kindly
Therefore it is desirable to characterize the available progestins provided by Dr. D.P. McDonnell as described previously [3].
according to their potential for development into illicit designer
androgens.
While previous studies have systematically examined andro- 2.3. Yeast culture
gen receptor (AR) binding for a variety of natural and synthetic
androgens and related steroids, such binding assays do not eval- Yeast transformants were grown overnight at 30 ◦ C with vig-
uate post-binding biological activity nor can they distinguish orous orbital shaking at 300 rpm in CSM-leu-ura (AR, BIO101)
whether a steroid binding to the AR is an agonist or antagonist. or CSM-trp-ura (PR, BIO101). Following overnight culture, the
Measurement of receptor interaction by yeast assay has proven yeast culture was subcultured in fresh medium and allowed to
robust and reproducible in a wide variety of pharmacological grow until early mid-log phase (OD600 ∼1.0).
and toxicological contexts and is a rapid and inexpensive first
pass screen to determine relative relationships between struc- 2.4. Yeast-based AR and PR bioassays
ture and bioactivity [15–20]. Yeast lack homologous steroid
hormone receptors yet contain enough conserved basal tran- For the AR bioassay, yeast from mid-log phase growth was
scriptional machinery to allow a recombinant human receptor diluted to OD600 = 1.0 in selective medium (CSM-leu-ura) plus
to facilitate the transcription of a co-transformed reporter vector 100 ␮M CuSO4 to induce receptor production via the CUP1
(␤-galactosidase) linked to a steroid responsive promoter [21]. promoter. For the PR bioassay, yeast was diluted to OD600 = 0.7
In this study, we have used yeast-based in vitro androgen in selective medium (CSM-trp-ura). Diluted yeast cultures
receptor and progesterone receptor (PR) bioassays to conduct were aliquoted into 24-well culture plates (500 ␮l/well) and
a systematic structure–activity study of a comprehensive set then treated with steroid concentrations ranging from 10−11 M
of commercially available progestins and related structures, to to10−2 M (in a final volume of 5 ␮l). Following incubation yeast
identify those compounds or structures with the highest intrinsic was lysed and assayed for ␤-galactosidase activity as previously
androgenic potency and greatest potential as a synthesis starting described [3].
L. McRobb et al. / Journal of Steroid Biochemistry & Molecular Biology 110 (2008) 39–47 41

2.5. Standard curves and determination of EC50 values

A testosterone or progesterone standard dose–response curve


was included in triplicate in every AR and PR bioassay,
respectively. AR and PR standard curves were calculated by
transforming the raw enzymatic read-out to a percentage of max-
imal activity of testosterone (AR assay) or progesterone (PR
assay) and fitted to a 4-parameter sigmoidal curve with variable
slope using Sigmaplot version 8.0. The EC50 values from the
fitted curves for each test steroid were used to determine the
relative potency of the test steroid with respect to testosterone
(AR bioactivity) or progesterone (PR bioactivity). AR bioac-
tivity of steroid relative to testosterone was determined using
the equation: 100 × EC50 [steroid]/EC50 [T]. PR bioactivity of
steroid relative to progesterone was determined using the follow-
ing: 100 × EC50 [steroid]/EC50 [P]. AR/PR potency ratio was
determined by dividing AR bioactivity by PR bioactivity.

3. Results

A total of 45 commercially available progestins and related


steroids were evaluated using a yeast bioassay for their AR and
PR bioactivity. The mean EC50 , relative bioactivity and AR/PR
potency ratio were determined for each steroid and are shown in
Table 1 (progesterone derivatives) and Table 2 (testosterone and
19-nortestosterone derivatives). Representative dose–response
curves for both the AR and PR bioassays are shown in Fig. 1A
and B, respectively. All steroids showed full agonist activity in
the AR bioassay, reaching close to 100% of maximal T apart
from nilutamide which was a partial agonist, reaching a plateau
of 30% of maximal bioactivity (not shown). Significant deviation
from maximum values in the PR bioassay was seen only with
the PR antagonist, mifepristone (53% of maximal, Fig. 1B) and
the prodrug, ethylestrenol (63% of maximal, not shown). Fig. 1. Representative dose–response curves for various steroid classes com-
paring androgen (A) and progestin (B) bioactivity levels in the yeast bioassays.
3.1. AR bioassay Yeast cultures were treated with serial dilutions of each steroid: (䊉) proges-
terone; () testosterone; () altrenogest (A) or 19-nortestosterone (B); ()
17␣-hydroxyprogesterone; () 19-norprogesterone; () mifepristone. Standard
The mean EC50 for testosterone was 4.4 ± 0.47 nM (n = 45) curves (dotted lines) for testosterone (AR bioassay) and progesterone (PR bioas-
in the AR bioassay (AR bioactivity of 100). A compari- say) were included in each experiment and relative activity was calculated as
son of the AR bioactivity of synthetic progestins and related a percentage of maximal activity of each in the respective assays. Each point
steroids in the androgen bioassay ranged from 0.006 for the 19- represents the mean of at least three independent experiments ±S.E.M.
norprogesterone derivative nestorone (Table 1) to 688 for the
19-nortestosterone derivative altrenogest (Table 2). In general, backbone (17␣-hydroxyprogesterone) decreased AR bioactiv-
testosterone and 19-nortestosterone derivatives, with the typi- ity from 500- to 0.21-fold. 17␣-Hydroxyprogesterone (OHP),
cal 3-keto-4-ene structure of the steroid nucleus and a hydroxyl a naturally occurring precursor of bioactive steroids, is inactive
substitution at the 17␤-position (17␤-OH), showed particularly with respect to AR and PR binding unless esterified [22,23].
strong AR activation (Table 2, Fig. 1A). Removal or substitu- Commonly marketed OHP derivatives included in this study
tion of the 3-keto group or replacement of the 4-ene structure (chlormadinone, cyproterone, cyproterone acetate and delmadi-
with a 5-10-ene (such as those steroids labeled “prodrugs” in none acetate) which have a 6-chloro-6-ene substitution, with
Table 2) resulted in mid-range AR bioactivity that was reduced or without acetylation of the 17␣-hydroxyl group, increased
between 10- and 100-fold compared to T. Substitution of the AR bioactivity relative to OHP but were still weak agonists
17␤-position with an acetyl group led to substantially reduced in this yeast assay (0.45–2.5). These steroids are reportedly
AR bioactivity as seen for the progesterone, 19-norprogesterone potent anti-androgens in vivo [24], while weaker androgen
and 17␣-hydroxyprogesterone derivatives (Table 1, Fig. 1A). antagonism is also a feature of megestrol acetate, a 6-methyl-
With an AR bioactivity of 10, progesterone showed the 6-ene derivative [7] which in this assay showed a potency
highest androgenic potency compared to its related deriva- value of 3.3. The non-steroidal antagonists, hydroxyflutamide
tives. Addition of a 17␣-hydroxy group to the progesterone and nilutamide, also showed transactivation in this yeast assay
42 L. McRobb et al. / Journal of Steroid Biochemistry & Molecular Biology 110 (2008) 39–47

Table 1
Relative potency of progesterone derivatives on the human androgen and progesterone receptors in yeast

Steroid CAS number Androgen receptor bioactivity Progesterone receptor bioactivity AR/PR potency ratiod

EC50 (nM)a Relative potencyb EC50 (nM)a Relative potencyc

Progesterone derivatives
Progesterone 57-83-0 46 9.6 3.9 100 0.096
Dydrogesterone 152-62-5 572 0.77 7.4 53 0.015
17␣-Hydroxyprogesterone derivatives
17␣-Hydroxyprogesterone 68-96-2 2137 0.21 1221 0.32 0.7
Medroxyprogesterone acetate 71-58-9 92 4.8 5.4 72 0.07
Megestrol acetate 595-33-5 132 3.3 4.8 81 0.04
Chlormadinone 1961-77-9 177 2.5 5.1 76 0.03
Cyproterone acetate 427-51-0 431 1.0 52 7.5 0.01
Cyproterone 2098-66-0 925 0.48 59 6.6 0.007
Delmadinone acetate 13698-49-2 972 0.45 39 10 0.045
19-Norprogesterone derivatives
Norprogesterone 472-54-8 115 3.8 2.2 177 0.02
Gestonorone caproate 1253-28-7 131 3.4 12 33 0.1
ZK 187 226 264186-52-9 228 1.9 2.5 156 0.01
Promegestone (R5020) 34184-77-5 1197 0.37 1.8 216 0.002
Org 2058 24320-06-7 4435 0.10 2.4 163 0.0006
Nestorone 7759-35-5 77900 0.006 1.6 244 0.00002
a EC50 (concentration yielding half of the maximal effect) determined from sigmoidal dose response curves.
b Bioactivity of steroid relative to testosterone (100 × EC50 [steroid]/EC50 [T]).
c Bioactivity of steroid relative to progesterone (100 × EC50 [steroid]/EC50 [P]).
d AR potency divided by PR potency.

although their AR bioactivity was reduced over 40,000-fold 3.2. PR bioassay


compared to testosterone (EC50 ’s ∼17,000 nM, AR potency
of 0.026; not shown). Similarly, the two 19-nortestosterone The EC50 of progesterone was 3.9 ± 0.39 nM (n = 38) in the
derivatives which show AR antagonism [25,26] exhibited very PR bioassay (Table 1, Fig. 1B). The results in Table 1 show
little AR bioactivity (dienogest, 0.15; mifepristone, 0.015). The that there was less distinct division of activity between the 17␣-
19-norprogesterone derivatives also showed very weak ago- hydroxyprogesterone and 19-nortestosterone structural classes
nist activity with relative potencies in the range of 0.006– of progestins for PR, compared to AR, bioactivity. The most
3.8 (Table 1). potent progestins included both 19-norprogesterone and 19-
Although 17␤-substitution reduced AR bioactivity in the nortestosterone derivatives. The 19-norprogesterone derivatives
above steroids, 17␣-substitution on the testosterone, 19- showed the most consistent level of PR bioactivity with six out of
nortestosterone, or 18-methyl-19-nortestosterone backbone or seven steroids tested having potencies greater than progesterone.
in combination with the 4,9,11-triene structure resulted in a While the most potent progestin was found amongst the 19-
series of steroids with strong AR bioactivity (Table 2). Fig. 2 nortestosterone derivatives, the weakest progestin also lay within
shows the chemical structures of the 17␣-alkylated testosterone this structural group. Altrenogest showed a 13-fold greater activ-
and 19-nortestosterone derivatives and their associated AR and ity than progesterone in this bioassay while mifepristone, a
PR potencies to allow easy comparison of trends between the PR antagonist, showed 500-fold decreased activity (Table 2,
groups. Within those shown in Fig. 2 the commercially marketed Fig. 1B). The 17␣-hydroxyprogesterone derivatives, although
progestins include the 17␣-ethynylated derivatives, ethisterone, antagonistic with regard to AR bioactivity, remain PR agonists
gestrinone, norgestrel and norethisterone. Ethynylation resulted and in this assay with values in the range of 6.6–8.1 (Table 1).
in decreased AR bioactivity in all steroid backbones (potency In the PR bioassay, the classical androgens T (0.2) and DHT
values from 19 to 49). Other short 17␣-alkyl substitutions (0.7) showed little bioactivity whereas removal of the 19-methyl
however showed increased AR bioactivity in comparison. AR group (as in 19-nortestosterone/nandrolone) increased PR bioac-
bioactivity for structures with 17␣-methyl substitutions ranged tivity 50-fold relative to T. Many commercially available
from 40 (17␣-methyltestosterone) to 244 (methyltrienolone), progestins are ethynylated derivatives of 19-nortestosterone.
while structures with 17␣-ethyl substitutions included THG In contrast to the effect on AR bioactivity, ethynylation of
(102), norbolethone (222) and norethandrolone (105). The the 17␣-position increased PR bioactivity as seen between
most potent androgen tested in this study, altrenogest, had T (0.2) and ethisterone (17) or 19-nortestosterone (12) and
a 17␣-allyl substitution in the 4,9,11-triene backbone. The norethisterone (80) (Table 2, Fig. 2). Most of the 17␣-
potency of this substitution was not maintained however in the alkylated 19-nortestosterone derivatives with the alkyl side
4-ene derivatives of 17␣-allylnandrolone (16) and 17␣-allyl-18- chains (methyl, ethyl and allyl) in the 4-ene backgrounds showed
methylnandrolone (19) (Fig. 2). an increase in PR bioactivity (range of 28–122) compared to
L. McRobb et al. / Journal of Steroid Biochemistry & Molecular Biology 110 (2008) 39–47 43

Table 2
Relative potency of testosterone and 19-nortestosterone derivatives on the human androgen and progesterone receptors in yeast
Steroid CAS number Androgen receptor Progesterone receptor AR/PR potency Relevant
bioactivity bioactivity ratiod substitution
EC50 (nM)a Relative EC50 (nM)a Relative 13␤ 17␣
potencyb potencyc

C19 methyl
Testosterone 58-22-0 4.4 100 1565 0.2 500 –CH3 –
Ethisterone 434-03-7 23.1 21 23 17 1.2 –CH
17-Methyltestosterone 58-18-4 11.0 40 175 2.2 18.2 –CH3
C19 nor
Nandrolone (19-nortestosterone) 434-22-0 4.3 102 34 12 8.5 –CH3 –
Norethisterone 68-22-4 22.6 19 4.9 80 0.24 –CH
17␣-Methylnandrolone 514-61-4 5.6 79 14 28 2.8 –CH3
Norethandrolone 52-78-8 4.2 105 4.2 93 1.1 –C2 H5
17␣-Allylnandrolonee 7538-46-5 27.4 16 8.1 48 0.33 –C3 H5
C19 nor
18-Methylnandrolonee 793-55-5 3.3 133 60 6.5 20.5 –C2 H5 –
Norgestrel 6533-00-2 9 49 2.6 150 0.33 –CH
17␣,18-Dimethylnandrolonee 14115-37-8 6.5 68 3.4 115 5.9 –CH3
Norbolethonee 1235-15-0 2.0 222 3.2 122 1.8 –C2 H5
17␣-Allyl-18-methylnandrolonee 141300-01-8 22.5 19 4.9 80 0.24 –C3 H5
4,9,11-Triene
Trenbolone 10161-33-8 14.3 31 802 0.5 62 –CH3 –
Gestrinone (R2323) 16320-04-0 18.9 23 49 8.0 2.9 –C2 H5 –CH
Methyltrienolone (R1881) 965-93-5 1.8 244 0.9 433 0.56 –CH3 –CH3
Tetrahydrogestrinone (THG)f 618903-56-3 4.3 102 4.1 95 0.34 –C2 H5 –C2 H5
Altrenogest 850-52-2 0.64 688 0.3 1300 0.53 –CH3 –C3 H5
Prodrugs (lack 3-keto or 4-ene)
Ethylestrenol 965-90-2 37.9 12 56 7.0 1.7
Norethynodrel 68-23-5 40.0 11 25 16 0.7
Tibolone 5630-53-5 70.0 6.3 422 0.9 7
Lynestrenol 52-76-6 86.0 5.1 95 4.1 1.2
Ethynodiol 1231-93-2 284 1.5 208 1.9 0.8
Allylestrenol 432-60-0 663 0.7 381 1.0 0.7
Norgestimate 35189-28-7 2496 0.18 11 35 0.005
Miscellaneous
Dihydrotestosterone (DHT) 521-18-6 2.8 157 546 0.7 224
Danazol 17230-88-5 7.8 59 645 0.6 98
3-Keto-desogestrel 54058-10-1 25.0 18 1.2 325 5.5
Steroidal AR antagonists
Dienogest 65928-58-7 2957 0.15 317 1.2 0.13
Mifepristone 84371-65-3 30430 0.015 2243 0.17 0.09
a EC50 (concentration yielding half of the maximal effect) determined from sigmoidal dose response curves.
b Bioactivity of steroid relative to testosterone (100 × EC50 [steroid]/EC50 [T]).
c Bioactivity of steroid relative to progesterone (100 × EC50 [steroid]/EC50 [P]).
d AR potency divided by PR potency.
e Synthesized at the School of Chemistry, University of Sydney, Australia.
f Synthesized at the National Measurement Institute, Sydney, Australia.

the equivalent structures without 17␣-substitutions (0.5–12), increased potency is consistent with the decreased dissociation
however there was either little change or decreased PR bioac- rate after extended incubation in receptor binding studies found
tivity compared to their ethynylated derivatives (80–150). In the for this substitution [27]. Compounds that dissociate slower tend
4,9,11-triene background, however, the PR bioactivity was sig- to be more potent agonists in vivo as they can sustain a response
nificantly increased as seen for example between gestrinone (8) in the nucleus.
and its three alkylated congeners, methyltrienolone (433), THG
(95) and altrenogest (1300). PR bioactivity was also sensitive 3.3. Comparison of AR and PR bioactivity
in these structural backgrounds to the replacement of the C13
methyl with a C18 methyl group (Table 2, Fig. 2) which in gen- Fig. 3 plots the AR and PR bioactivity of each steroid.
eral increased both PR and AR bioactivity in this assay. This Deviations from the line reflecting proportionate bioactivity are
44 L. McRobb et al. / Journal of Steroid Biochemistry & Molecular Biology 110 (2008) 39–47

Fig. 2. Chemical structures of 17␣-alkylated testosterone, 19-nortestosterone and 18-methyl-19-nortestosterone derivatives used in this study with their associated
AR and PR potency values as determined as described in Section 2. Structures are grouped in columns according to their basic skeletal structure: (A) C19 methyl,
(B) 4,9,11-triene, (C) C19 nor with C13 ethyl and (D) C19 nor with C18 methyl; in rows according to their 17␣-substitution.

seen mainly with testosterone derivatives possessing dispropor- AR/PR potency ratios for the former group which include
tionately high-androgenic activity whereas 19-norporgesterone those structures unencumbered at the 17␣-position, such as
derivatives had dissociation of high progestin with low- T, nandrolone and trenbolone (Table 2). The 17␣-alkylated
androgenic bioactivity. This is also reflected in the very high 19-nortestosterone derivatives with both potent androgen and
progestin bioactivities showed more proportionate activities
in Fig. 3 but also a subsequent decrease in the AR/PR ratio
(Table 2), not due to decreases in androgenicity per se but sig-
nificant increases in PR bioactivity.

4. Discussion

In this study, we used in vitro yeast bioassays as a rapid and


inexpensive first pass screen to determine the intrinsic andro-
genic potential of a comprehensive set of commercially available
progestins and related steroid structures. The most potent andro-
gens determined by the yeast bioassay, and hence those most
at risk for abuse in a sports doping context were derived from
Fig. 3. Comparison of AR vs. PR potency of structural groups tested in this either testosterone or 19-nortestosterone and maintained the
study. Mean values of AR potency are plotted vs. mean PR potency for each basic steroidal structure of the natural androgen, T, with a 3-keto
group of structurally related steroids with bidirectional error bars (S.E.M.). The group and a 17␤-OH substitution. This finding is consistent with
potency values are relative to testosterone (1 0 0) in the androgen bioassay (AR) previous studies showing that substitutions at the C3 or C17 posi-
and progesterone (1 0 0) in the progestin bioassay (PR). Open symbols represent
tion of the steroid nucleus are important for androgen receptor
subsets of the testosterone and 19-nortestosterone derivatives as follows: () AR
antagonists, () prodrugs, () androgens, () 17␣-alkylated derivatives and (♦) binding [19,22,28,29].
4,9,11-trienes. Closed symbols represent derivatives of (䊉) 19-norprogesterone, We previously demonstrated using this yeast bioassay that
() progesterone (P) and 17␣-hydroxyprogesterone (OHP). THG, with its 17␣-ethyl substitution is a potent androgen [3],
L. McRobb et al. / Journal of Steroid Biochemistry & Molecular Biology 110 (2008) 39–47 45

a finding since confirmed by others using micro-array analysis genicity, the ability in this study to synthesize norbolethone and
showing that THG carries the signature of a potent androgen the two allyl containing steroids from norgestrel by relatively
[4]. We now extend this finding to observe that substitution simple substitution of the ethynyl group of norgestrel, and the
at the 17␣-position with alkyl side chains in a testosterone or previous identification of THG [2] which is derived from gestri-
19-nortestosterone skeleton or in the presence of three conju- none by hydrogenation of the 17␣-ethynyl side chain, shows that
gated double bonds (the 4,9,11-trienes) resulted in a subset of these ethynylated compounds have more potential as lead com-
molecules with very potent androgenic and progestogenic activ- pounds for the production of either more androgenic compounds
ity (Table 2). Commercially marketed progestins in this subset or compounds with an equivalent androgenicity but novel struc-
include the 17␣-ethynylated derivatives, ethisterone, norethis- tures that may be undetectable by current screening methods.
terone, norgestrel and the 4,9,11-triene, gestrinone. Their AR The 17␣-ethyl and allyl-substituted derivatives, with or without
bioactivity was in the range of 19–23 although gestrinone is 18-methyl substitutions, of the ethynylated testosterone deriva-
reported to have high binding affinity in receptor binding stud- tive, ethisterone, were not considered in this study but may also
ies and androgenic effects on the rat ventral prostate although pose a similar risk. Although previous studies have shown that
at 10-fold higher concentrations than T [30]. Both norethis- methyl-substituted 4,9-dienes have lower receptor binding affin-
terone and norgestrel have equivalent if not greater androgenicity ity for AR than mono- or tri-enes [27], these compounds still
than gestrinone in these same studies [27,30]. This suggests that appear to have intrinsic androgenicity and further 17␣ conver-
they represent likely starting materials for novel illicit designer sion may create more androgenic but undetectable compounds
androgens. that should be further considered for their associated risk.
Synthetic modification of the 17␣-ethynyl side chains Progestins classified as progesterone, 19-norprogesterone
of gestrinone, norethisterone and norgestrel, can form and 17␣-hydroxyprogesterone derivatives displayed little andro-
either methyl (methyltrienolone, 17␣-methylnandrolone or genic activity in the yeast AR bioassay. This suggests that they
17␣,18-dimethylnandrolone) or ethyl (THG, norbolethone and possess low risk as sports doping agents. Loss of the 17␤-OH
norethandrolone) substituents, respectively. Saturation of this group in these molecules results in the disruption of hydro-
side chain significantly increased AR bioactivity between 1.5- gen bonding while replacement with a bulky acetyl group is
and 10-fold for the methyl-substituted structure and 4- to 5-fold thought to reduce binding due to an increased steric hindrance
for the ethyl-substituted structure but had less significant effects in the hAR binding pocket [28,38–41]. The C16 substitution and
on PR bioactivity. Further increases in potency in vivo would be even bulkier substitutions found in the 17␤-position of the 19-
expected of these 17␣-alkylated derivatives due to the fact that norprogesterone derivatives tested in this study would provide
metabolism to the inactive 17-keto derivative has been shown even further steric hindrance in the binding pocket of the AR
to be eliminated by the alkylation at the 17␣-position, result- and may explain their very low level of AR bioactivity deter-
ing in an increased half-life and potency [31,32]. Indeed these mined in this assay. Among the 19-norprogesterone derivatives
compounds were originally developed in the 1950s to increase are the newer or fourth generation of progestins designed to
oral bioavailability, however their clinical suitability was limited further increase relative progestational potency while reducing
by an associated increase in hepatotoxicity [33–35]. The most androgenicity [9,42]. In the PR bioassay, five of the six steroids
potent progestin and androgen in this study however was the tested in this group were more potent than progesterone while
17␣-allyl-substituted 4,9,11-triene, altrenogest, a potent veteri- corresponding androgenic activity was in the range of minimal
nary progestin [36,37] which was 6 times more potent as an activation observed with androgen antagonists.
androgen than testosterone and 13 times more potent a pro- In general the yeast assay showed good consistency with
gestin than progesterone. However, when this side chain was published androgen receptor binding studies resulting in a
examined in the 4-ene skeleton of 17␣-allylnandrolone and 17␣- similar ranking of steroids for progestin and androgenic activ-
allyl-18-methylnandrolone, the androgenic potential decreased ities [7,19,30,43]. One notable discrepancy however was that
in comparison to the ethynylated derivatives. It may be deduced medroxyprogesterone acetate (MPA) showed less AR bioactiv-
from these results that short, saturated side chain substitutions ity than predicted from previous binding studies. In the latter,
increase AR binding and activation while longer chain lengths MPA had equivalent binding activity to the more androgenic
reduce receptor fit. This is consistent with AR crystal structure progestins, norgestrel and norethisterone, while in our bioassay
studies which show that longer side chains in the 17␣-position MPA had much lower bioactivity compared to the latter two pro-
appear to interfere with the hydrogen bonding between the 17␤- gestins. This result highlights the essential difference between
OH group and the receptor [38]. Studies have shown the greater receptor binding studies and bioassays where the latter requires
inherent flexibility of the unsaturated triene backbone [27] which not only receptor binding but subsequent transactivation as well.
may allow better receptor fit in the presence of longer side chains Results in our in vitro bioassay are more consistent with the in
and may also explain the increase in cross-reactivity between vivo assay on rat prostate which showed androgenic activity but
steroid receptors of this group. only at very high doses of this steroid [30,44]. It is however
Gestrinone is the only marketed progestin currently on the a limitation of in vitro bioassays that they do not evaluate in
WADA banned list but the results from this study and others vivo activation or inactivation of the parent steroid to metabo-
[27,30] show that the other ethynylated steroids have intrinsic lites with or without, respectively, bioactivity [45,46]. In this
androgenicity equal to or greater than gestrinone and present as study, the relatively low activity of the so-called prodrugs, such
much risk for potential abuse. Apart from their intrinsic andro- as lynestrenol and norethyndrel, which lack intrinsic activity
46 L. McRobb et al. / Journal of Steroid Biochemistry & Molecular Biology 110 (2008) 39–47

without further metabolism to the active norethisterone [47], [7] A.E. Schindler, C. Campagnoli, R. Druckmann, J. Huber, J.R. Pasqualini,
illustrate the limited metabolic capabilities of yeast with respect K.W. Schweppe, J.H. Thijssen, Classification and pharmacology of pro-
to steroid metabolism [15]. gestins, Maturitas 46 (Suppl. 1) (2003) S7–S16.
[8] C. Djerassi, L. Miramontes, G. Rosenkranz, F. Sondheimer, L.I.V.
While in vitro assays are no replacement for subsequent Steroids, Synthesis of 19-nor-17alpha-ethynyltestosterone and 19-nor-
in vivo testing, the yeast assay used in this study provides a 17alpha-methyltestosterone, J. Am. Chem. Soc. 76 (1954) 4092–4094.
rapid and inexpensive first pass screen for determining the rela- [9] S. Rowlands, Newer progestogens, J. Fam. Plann. Reprod. Health Care 29
tionship between structure and androgenicity. Analysis of a (1) (2003) 13–16.
comprehensive list of steroids in this study has given thresh- [10] R.M. Evans, The steroid and thyroid hormone receptor superfamily, Sci-
ence 240 (4854) (1988) 889–895.
olds within which to examine and classify new structures to [11] A.O. Brinkmann, P.W. Faber, H.C. van Rooij, G.G. Kuiper, C. Ris, P.
determine the potential for in vivo androgenic (and progesto- Klaassen, J.A. van der Korput, M.M. Voorhorst, J.H. van Laar, E. Mulder,
genic) activity. This has also allowed us to predict from current et al., The human androgen receptor: domain structure, genomic organi-
marketed progestins which of those structures are at greatest zation and regulation of expression, J. Steroid Biochem. 34 (1–6) (1989)
risk of abuse or as a synthesis starting point for further mod- 307–310.
[12] A.O. Brinkmann, J. Trapman, Genetic analysis of androgen receptors in
ification. We have shown that while 17␣-hydroxyprogesterone development and disease, Adv. Pharmacol. 47 (2000) 317–341.
and 19-norprogesterone derivatives have relatively weak intrin- [13] WADA, The World Anti-Doping Code—the 2007 Prohibited List,
sic AR bioactivity and represent little risk of abuse in a sports www.wada-ama.org/rtecontent/document/2007 List En.pdf, 2007
doping context, we have identified strong AR agonists among (accessed May 5, 2007).
the 17␣-alkylated-19-nortestosterone derivatives. The intrinsic [14] E. Sakiz, G. Azadian-Boulanger, R2323—an original contraceptive com-
pound, Excerpta Med. 219 (1971) 865–971.
androgenicity of the 17␣-ethynylated progestins that are com- [15] K.W. Gaido, L.S. Leonard, S. Lovell, J.C. Gould, D. Babai, C.J. Portier, D.P.
mercially marketed rank them as high risk for use as sports McDonnell, Evaluation of chemicals with endocrine modulating activity in
doping agents, and their relatively simple conversion to more a yeast-based steroid hormone receptor gene transcription assay, Toxicol.
potent androgens makes them high risk as lead compounds for Appl. Pharmacol. 143 (1) (1997) 205–212.
the synthesis of undetectable designer androgens. Consideration [16] P. Roy, S. Franks, M. Read, I.T. Huhtaniemi, Determination of androgen
bioactivity in human serum samples using a recombinant cell based in vitro
should be given to their addition to the WADA banned substances bioassay, J. Steroid Biochem. Mol. Biol. 101 (1) (2006) 68–77.
list and priority given to the development of GC–MS methods [17] S.F. Arnold, M.K. Robinson, A.C. Notides, L.J. Guillette Jr., J.A. McLach-
for their potential derivatives. lan, A yeast estrogen screen for examining the relative exposure of cells
to natural and xenoestrogens, Environ. Health Perspect. 104 (5) (1996)
544–548.
Acknowledgements
[18] T.F. Bovee, R.J. Helsdingen, P.D. Koks, H.A. Kuiper, R.L. Hoogenboom,
J. Keijer, Development of a rapid yeast estrogen bioassay, based on the
This work was supported by funding from the World Anti- expression of green fluorescent protein, Gene 325 (2004) 187–200.
Doping Agency and the Australian Anti-Doping Research Panel. [19] H. Fang, W. Tong, W.S. Branham, C.L. Moland, S.L. Dial, H. Hong, Q.
L. McRobb was co-funded by a University Postgraduate Award Xie, R. Perkins, W. Owens, D.M. Sheehan, Study of 202 natural, synthetic,
and environmental chemicals for binding to the androgen receptor, Chem.
(UPA), the Department of Medicine, University of Sydney and
Res. Toxicol. 16 (10) (2003) 1338–1358.
the Heart Research Institute, Sydney. The authors are grateful [20] A.K. Death, K.C. McGrath, D.J. Handelsman, Valproate is an anti-androgen
to Dr. P. Stamford and Mr. S. Wilkinson, School of Chemistry, and anti-progestin, Steroids 70 (14) (2005) 946–953.
University of Sydney for help with steroid synthesis, and to Dr. [21] D. Metzger, J.H. White, P. Chambon, The human oestrogen receptor func-
R. Sitruk-Ware of the Population Council, New York, for the tions in yeast, Nature 334 (6177) (1988) 31–36.
[22] J.L. McGuire, C.D. Bariso, A.P. Shroff, Interaction between steroids and a
supply of Nesterone.
uterine progestogen specific binding macromolecule, Biochemistry 13 (2)
(1974) 319–322.
References [23] H.E. Smith, R.G. Smith, D.O. Toft, J.R. Neergaard, E.P. Burrows, B.W.
O’Malley, Binding of steroids to progesterone receptor proteins in chick
[1] W. Kondro, Athletes’ “designer steroid” leads to widening scandal, Lancet oviduct and human uterus, J. Biol. Chem. 249 (18) (1974) 5924–5932.
362 (9394) (2003) 1466. [24] D. Raudrant, T. Rabe, Progestogens with antiandrogenic properties, Drugs
[2] D.H. Catlin, M.H. Sekera, B.D. Ahrens, B. Starcevic, Y.C. Chang, C.K. 63 (5) (2003) 463–492.
Hatton, Tetrahydrogestrinone: discovery, synthesis, and detection in urine, [25] R.H. Foster, M.I. Wilde, Dienogest, Drugs 56 (5) (1998) 825–835.
Rapid Commun. Mass Spectrom. 18 (12) (2004) 1245–1249. [26] R. Sitruk-Ware, I.M. Spitz, Pharmacological properties of mifepristone:
[3] A.K. Death, K.C.Y. McGrath, R. Kazlauskas, D.J. Handelsman, Tetrahy- toxicology and safety in animal and human studies, Contraception 68 (6)
drogestrinone (THG) is a potent androgen and progestin, J. Clin. (2003) 409–420.
Endocrinol. Metab. 89 (5) (2004) 2498–2500. [27] J. Delettre, J.P. Mornon, T. Ojasoo, J.P. Raynaud, An approach to the map-
[4] F. Labrie, V. Luu-The, E. Calvo, C. Martel, J. Cloutier, S. Gauthier, ping of the steroid hormone receptor, in: F. Bresciani (Ed.), Perspectives in
P. Belleau, J. Morissette, M.H. Levesque, C. Labrie, Tetrahydrogestri- Steroid Receptor Research, Raven Press, New York, 1980, pp. 1–21.
none induces a genomic signature typical of a potent anabolic steroid, J. [28] N. Poujol, J.M. Wurtz, B. Tahiri, S. Lumbroso, J.C. Nicolas, D. Moras,
Endocrinol. 184 (2) (2005) 427–433. C. Sultan, Specific recognition of androgens by their nuclear recep-
[5] M.H. Sekera, B.D. Ahrens, Y.C. Chang, B. Starcevic, C. Georgakopoulos, tor. A structure–function study, J. Biol. Chem. 275 (31) (2000) 24022–
D.H. Catlin, Another designer steroid: discovery, synthesis, and detec- 24031.
tion of ‘madol’ in urine, Rapid Commun. Mass Spectrom. 19 (6) (2005) [29] T. Ojasoo, J.P. Raynaud, Unique steroid congeners for receptor studies,
781–784. Cancer Res. 38 (11) (1978) 4186–4198.
[6] D.H. Catlin, B.D. Ahrens, Y. Kucherova, Detection of norbolethone, an [30] J.P. Raynaud, M.M. Bouton, M. Moguilewsky, T. Ojasoo, D. Philibert, G.
anabolic steroid never marketed, in athletes’ urine, Rapid Commun. Mass Beck, F. Labrie, J.P. Mornon, Steroid hormone receptors and pharmacology,
Spectrom. 16 (13) (2002) 1273–1275. J. Steroid Biochem. 12 (1980) 143–157.
L. McRobb et al. / Journal of Steroid Biochemistry & Molecular Biology 110 (2008) 39–47 47

[31] Y. Shinohara, S. Baba, Stable isotope methodology in the pharmacokinetic Crystallographic structures of the ligand-binding domains of the andro-
studies of androgenic steroids in humans, Steroids 55 (4) (1990) 170–176. gen receptor and its T877A mutant complexed with the natural agonist
[32] R.V. Quincey, C.H. Gray, The metabolism of [1,2-3H]17-alpha- dihydrotestosterone, Proc. Natl. Acad. Sci. U.S.A. 98 (9) (2001) 4904–
methyltestosterone in human subjects, J. Endocrinol. 37 (1) (1967) 37–55. 4909.
[33] K.G. Ishak, H.J. Zimmerman, Hepatotoxic effects of the [41] C.A. Marhefka, B.M. Moore 2nd, T.C. Bishop, L. Kirkovsky, A. Mukherjee,
anabolic/androgenic steroids, Semin. Liver Dis. 7 (3) (1987) 230–236. J.T. Dalton, D.D. Miller, Homology modeling using multiple molecular
[34] A. Cabasso, Peliosis hepatis in a young adult bodybuilder, Med. Sci. Sports dynamics simulations and docking studies of the human androgen receptor
Exerc. 26 (1) (1994) 2–4. ligand binding domain bound to testosterone and nonsteroidal ligands, J.
[35] V. Petera, K. Bobek, V. Lahn, Serum transaminase (GOT, GPT) and lactic Med. Chem. 44 (11) (2001) 1729–1740.
dehydrogenase activity during treatment with methyl testosterone, Clin. [42] R. Sitruk-Ware, New progestagens for contraceptive use, Hum. Reprod.
Chim. Acta 7 (1962) 604–606. Update 12 (2) (2006) 169–178.
[36] D.A. Redmer, B.N. Day, Ovarian activity and hormonal patterns in gilts [43] R. Sitruk-Ware, Pharmacological profile of progestins, Maturitas 47 (4)
fed allyl trenbolone, J. Anim. Sci. 53 (4) (1981) 1088–1094. (2004) 277–283.
[37] D. Hodgson, S. Howe, L. Jeffcott, S. Reid, D. Mellor, A. Higgins, Effect of [44] A. Phillips, D.W. Hahn, S. Klimek, J.L. McGuire, A comparison of the
prolonged use of altrenogest on behaviour in mares, Vet. J. 169 (3) (2005) potencies and activities of progestogens used in contraceptives, Contra-
322–325. ception 36 (2) (1987) 181–192.
[38] P.M. Matias, P. Donner, R. Coelho, M. Thomaz, C. Peixoto, S. Macedo, [45] N. Beresford, E.J. Routledge, C.A. Harris, J.P. Sumpter, Issues arising when
N. Otto, S. Joschko, P. Scholz, A. Wegg, S. Basler, M. Schafer, U. Egner, interpreting results from an in vitro assay for estrogenic activity, Toxicol.
M.A. Carrondo, Structural evidence for ligand specificity in the binding Appl. Pharmacol. 162 (1) (2000) 22–33.
domain of the human androgen receptor. Implications for pathogenic gene [46] D. Yin, H. Xu, Y. He, L.I. Kirkovsky, D.D. Miller, J.T. Dalton, Pharma-
mutations, J. Biol. Chem. 275 (34) (2000) 26164–26171. cology, pharmacokinetics, and metabolism of acetothiolutamide, a novel
[39] W. Bourguet, P. Germain, H. Gronemeyer, Nuclear receptor ligand-binding nonsteroidal agonist for the androgen receptor, J. Pharmacol. Exp. Ther.
domains: three-dimensional structures, molecular interactions and pharma- 304 (3) (2003) 1323–1333.
cological implications, Trends Pharmacol. Sci. 21 (10) (2000) 381–388. [47] F.Z. Stanczyk, Pharmacokinetics and potency of progestins used for hor-
[40] J.S. Sack, K.F. Kish, C. Wang, R.M. Attar, S.E. Kiefer, Y. An, G.Y. Wu, J.E. mone replacement therapy and contraception, Rev. Endocr. Metab. Disord.
Scheffler, M.E. Salvati, S.R. Krystek Jr., R. Weinmann, H.M. Einspahr, 3 (3) (2002) 211–224.

You might also like