ArticlePDF Available

Transformations of Steroids by Beauveria bassiana

Authors:

Abstract

The course of transformations of testosterone and its derivatives, including compounds with an additional C1,C2 double bond and/or a 17alpha-methyl group, a 17beta-acetyl group or without a 19-methyl group, by a Beauveria bassiana culture was investigated. The fungi promoted hydroxylation of these compounds at position 11alpha, oxidation of the 17beta-hydroxyl group, reduction of the C1,C2 or C4,C5 double bonds and degradation of the progesterone side-chain, leading to testosterone. The structure of 4-ene-3-oxo-steroids had no influence on regio- and stereochemistry of hydroxylation. In a similar manner, dehydroepiandrosterone was hydroxylated by Beauveria bassiana at position 11alpha, however, a small amount of 7alpha-hydroxylation product was also formed.
Transformations of Steroids by Beauveria bassiana
Ewa Huszcza
*
, Jadwiga Dmochowska-Gładysz, and Agnieszka Bartman
´ska
Department of Chemistry, Agricultural University, Norwida 25, 50-375 Wrocław, Poland.
Fax: 0048-071-3 283576. E-mail: huszcza@ozi.ar.wroc.pl
* Author for correspondence and reprint requests
Z. Naturforsch. 60 c, 103Ð108 (2005); received July 26/September 6, 2004
The course of transformations of testosterone and its derivatives, including compounds
with an additional C1,C2 double bond and/or a 17α-methyl group, a 17β-acetyl group or
without a 19-methyl group, by a Beauveria bassiana culture was investigated. The fungi pro-
moted hydroxylation of these compounds at position 11α, oxidation of the 17β-hydroxyl
group, reduction of the C1,C2 or C4,C5 double bonds and degradation of the progesterone
side-chain, leading to testosterone. The structure of 4-ene-3-oxo-steroids had no influence on
regio- and stereochemistry of hydroxylation. In a similar manner, dehydroepiandrosterone
was hydroxylated by Beauveria bassiana at position 11α, however, a small amount of 7α-
hydroxylation product was also formed.
Key words: Beauveria bassiana, Biotransformation, Steroids
Introduction
Beauveria, belonging to the Moniliaceae family
of fungi imperfecti, is a naturally occurring soil
fungus (Bidochka et al., 1998). It is a recognized
pathogen of more than 100 insect species (Hajek
and St. Leger, 1994), which has found an applica-
tion in agricultural biocontrol programs (Bing and
Lewis, 1991, 1992; Krueger and Roberts, 1997; Mu-
lock and Chandler, 2000).
The fungus Beauveria bassiana ATCC 7159
(also known as B. sulfurescens or Sporotrichum
sulfurescens) is one of the most frequently used
biocatalysts capable of performing reactions of a
different type, e.g. hydroxylation of saturated and
aromatic carbon atoms, keto-alcohol redox reac-
tion, alkene redox reaction, sulfide oxidation,
Baeyer-Villiger oxidation, glucosidation, epoxide
and ester hydrolysis and heteroatom dealkylation.
These results have been summarized in the review
article of Grogan and Holland (2000). The most
significant is the use of B. bassiana for selective
hydroxylation of a wide range of organic com-
pounds.
In contrast to many other fungi currently used
for biocatalysis, Beauveria has not been exten-
sively used for transformations of steroids. Previ-
ous research showed that B. bassiana promotes hy-
droxylation of 4-ene-3-oxo-steroids mainly at
position 11α(Griffiths et al., 1993; Bayunova et
al., 1989; C
ˇapek and Fassatiova, 1977; C
ˇapek et al.,
1966; Schubert et al., 1962) and, rarely, at positions
0939Ð5075/2005/0100Ð0103 $ 06.00 2005 Verlag der Zeitschrift für Naturforschung, Tübingen · http://www.znaturforsch.com ·
D
6βand 11α(Griffiths et al., 1993; C
ˇapek and Fas-
satiova, 1977; C
ˇapek et al., 1966). Hydroxylation
at positions 6α,11α,11βand 15αwas observed in
the B-norsteroid 17α,21-dihydroxy-B-nor-pregn-4-
en-3,20-dione (Sanada et al., 1977). Also the abil-
ity of B. bassiana to reduce the 17-ketone to a 17β-
alcohol (Bayunova et al., 1989), and to degrade the
progesterone side-chain leading to testosterone
(Schubert et al., 1962) was reported.
Therefore, it was of interest to us performing
comparative studies on various 4-ene-3-oxo-ste-
roids. We wanted to check whether the additional
C17 methyl group, the lack of the C19 methyl
group, and the additional C1,C2 double bond do
not alter the localisation of the hydroxylation pro-
cess.
Because the knowledge of biotransformations of
5-ene-steroids is much less documented compared
to 4-ene-3-oxo-steroids, we have chosen dehy-
droepiandrosterone (DHEA) as an additional
substrate for the tests with Beauveria bassiana.
There are no previous reports on 19-nortestoster-
one and dehydroepiandrosterone transformation
in a Beauveria bassiana culture.
Materials and Methods
Microorganism
Beauveria bassiana AM446 was obtained from
the Institute of Biology and Botany of the Medical
University of Wrocław. It was isolated from the
104 E. Huszcza et al. · Steroid Transformations by Beauveria bassiana
insect Pyrrhocoris apterus (Pyrrhocoridae) (ima-
go).
Conditions of cultivation and transformation
The fungi were incubated in 3% glucose and 1%
peptone, pH 5.9, and shaken at 27 C in 2 l Erlen-
meyer flasks with 300 ml of medium. After 3 d of
growth, 120 mg of a substrate, dissolved in 5 ml
of acetone or ethanol, were added, and the flasks
returned to the shaker. The products were ex-
tracted with chloroform after 3Ð10 d of trans-
formation (until the substrate was metabolized).
Product analysis
The composition of crude biotransformation
mixtures was analysed by TLC and GC. TLC was
carried out using silica-gel 60 plates (Merck) with
hexane/acetone (2:1 or 1:1 v/v) as eluent. Steroids
were detected by spraying the plates with H
2
SO
4
/
EtOH (1:1 v/v) followed by heating. Analytical
GC analysis was performed on a Hewlett Packard
5890A Series II GC instrument, using a HP-5 cap-
illary column (cross-linked 5% Ph-Me-Silicone,
30 m ¥0.53 mm ¥0.88 µm film thickness; temper-
ature program: 240 CÐ1 min, gradient 5 C/min
to 300 CÐ5 min). Biotransformation products
were separated by column chromatography using
silica gel 0.05Ð0.2 mesh (Merck) with a hexane/
acetone mixture (2:1 v/v) as eluent. Structures of
biotransformation products were determined on
the basis of
1
H NMR spectra, which were recorded
on a DRX 300 Bruker 300 MHz spectrometer in
Table I. Biotransformation of steroids by Beauveria bassiana.
Substrate Products Yield
a
(%)
Testosterone (1)11α-hydroxytestosterone (8) 55.4
5α-androstan-11α,17β-diol-3-one (9) 14.8
11α-hydroxyandrost-4-ene-3,17-dione (10) 10.5
5α-androstan-11α-ol-3,17-dione (11) 8.4
17α-Methyltestosterone (2)11α-hydroxy-17α-methyltestosterone (12) 95.3
19-Nortestosterone (3)11α-hydroxy-19-nortestoterone (13) 40.5
1-Dehydrotestosterone (4)11α-hydroxy-1-dehydrotestosterone (14) 56.4
11α-hydroxyandrost-1,4-diene-3,17-dione (15) 13.2
11α-hydroxytestosterone (8) 11.4
11α-hydroxyandrost-4-ene-3,17-dione (10) 9.3
1-Dehydro-17α-methyltestosterone (5)11α-hydroxy-1-dehydro-17α-methyltestosterone (16) 86.7
Progesterone (6)11α-hydroxytestosterone (8) 94.1
Dehydroepiandrosterone (7) 5-androsten-3β,11α,17β-triol (17) 59.6
7α-hydroxydehydroepiandrosterone (18) 13.1
androstenediol (19) 8.3
a
Yield determined by GC.
CDCl
3
, CD measurements, which were done on a
JASCO-715 spectropolarimeter in chloroform,
and optical rotation measurements, which were
performed on an AUTOPOL IV polarimeter in
acetone at 25 C.
Results and Discussion
In order to examine different structural factors
of a steroid on the biotransformation course, in-
cluding the presence of the C1,C2 double bond
and/or the additional 17α-methyl group and the
absence of either C19-methyl or C17-acetyl
groups, the following substrates have been chosen
for transformations by Beauveria bassiana AM446:
testosterone (1), 17α-methyltestosterone (2), 19-
nortestosterone (3), 1-dehydrotestosterone (4), 1-
dehydro-17α-methyltestosterone (5) and proges-
terone (6). As there is much less information
about hydroxylation of 5-androstenes compared to
4-ene-3-oxo-steroids, we also decided to explore
the bioconversion of dehydroepiandrosterone (7).
The fungus Beauveria bassiana was incubated
with the substrates until they were metabolised
(3Ð10 d). The results of the biotransformations
are presented by Fig. 1. The yield of products was
determined by GC analysis of the chloroform ex-
tract (Table I).
The structures of biotransformation products
were assigned mainly based on
1
H NMR spectra.
Both location and configuration of the newly in-
troduced hydroxyl group were determined by ana-
lysing differences between NMR spectra of the
E. Huszcza et al. · Steroid Transformations by Beauveria bassiana 105
O
OH
O
OH
O
OH
O
OH
O
OH
O
O
O
OH
OH
OH
OH
O
OH
OH
O
OH
OH
O
O
OH
O
O
OH
O
OH
OH
O
OH
OH
O
OH
OH
O
O
OH
O
OH
OH
O
O
OH
O
OH
OH
O
OH
OH
OH
OH OH
O
OH
+
+++
1
2
3
4
5
6
7
8910 11
12
13
14 15
16
17
+
+
810
8
++
18 19
Fig. 1. Metabolism of testosterone (1), 17α-methyltestosterone (2), 19-nortestosterone (3), 1-dehydrotestosterone
(4), 1-dehydro-17α-methyltestosterone (5), progesterone (6) and dehydroepiandrosterone (7)byBeauveria bassiana.
starting material and products (Table II, III), sup-
ported by literature data (Jones, 1973; Kirk et al.,
1990).
All the products obtained from 4-ene-3-oxo-ste-
roids transformations contained a 11α-hydroxyl
group, which was proved by a large downfield shift
of the 19-H
3
signal (but not 18-H
3
) and by the
broad multiplet profile for the 11β-H signal in the
region of δ3.86 ppm to 4.11 ppm as was reported
by Jones (1973) and Kirk et al. (1990). The spectral
106 E. Huszcza et al. · Steroid Transformations by Beauveria bassiana
Table II.
1
H NMR data for B. bassiana 4-ene-3-oxo-steroids transformation products and some substrates
a
.
Compound 4-H 17α-H 17α-CH
3
18-H 19-H 11β-H Other significant signals
85.72 3.69 t, J= 8.3 Hz Ð0.82 1.32 4.03 m Ð
9Ð3.69 t, J= 8.5 Hz Ð0.76 1.13 3.98 m Ð
10 5.73 ÐÐ0.93 1.32 4.05 m Ð
11 ÐÐ Ð0.89 1.15 4.02 m Ð
25.71 Ð1.20 0.90 1.19 ÐÐ
12 5.72 Ð1.23 0.92 1.32 4.05 m Ð
13 5.80 3.64 t, J= 8.5 Hz Ð0.78 ÐÐ Ð
13 5.81 3.68 t, J= 8.3 Hz Ð0.80 Ð3.86 m Ð
46.04 3.61 t, J= 8.4 Hz Ð0.79 1.21 Ð7.03 d, J= 10.2 Hz (1-H)
6.19 d, J= 10.1 Hz (2H)
14 6.06 3.63 t, J= 8.4 Hz Ð0.81 1.30 4.03 m 7.79 d, J= 10.3 Hz (1-H)
6.13 dd, J= 10.2 Hz, 2 Hz (2-H)
15 6.10 ÐÐ0.93 1.32 4.11 m 7.74 d, J= 10.3 Hz (1-H)
6.14 dd, J= 10.5 Hz, 2 Hz (2-H)
56.05 Ð1.18 0.92 1.24 Ð7.04 d, J= 10.2 Hz (1-H)
6.21 dd, J= 10.2 Hz, 2 Hz (2-H)
16 6.08 Ð1.20 0.93 1.32 4.09 m 7.79 d, J= 10.3 Hz (1-H)
6.13 dd, J= 10.2 Hz, 2 Hz (2-H)
a
Chemical shifts in ppm relative to Me
4
Si; solvent, CDCl
3
;J, coupling constant.
Table III.
1
H NMR data for dehydroepiandrosterone and its transformation products
a
.
Compound 6-H 3α-H CHOR 17α-H 18-H 19-H
75.34 d, J= 5.1 Hz 3.51 m, Wh = 23 Hz ÐÐ0.86 1.02
17 5.41 d, J= 5.8 Hz 3.53 m, Wh = 22 Hz 4.08 m (11β-H) 3.69 t, J= 8.5 Hz 0.78 1.16
18 5.62 d, J= 5.3 Hz 3.56 m, Wh = 22 Hz 3.95 m, Wh = 11 Hz (7β-H) Ð0.87 1.00
19 5.33 d, J= 5.2 Hz 3.50 m, Wh = 22 Hz Ð3.63 t, J= 8.5 Hz 0.75 1.01
a
Chemical shifts in ppm relative to Me
4
Si; solvent, CDCl
3
;J, coupling constant.
data of these compounds (Table II) correspond
very closely to those described in the literature for
11α-hydroxytestosterone (8) (Smith et al., 1990;
Kirk et al., 1990), 11α-hydroxyandrost-4-ene-3,17-
dione (10) (Kirk et al., 1990), 11α-hydroxy-17α-
methyltestosterone (12) (Huszcza and Dmochow-
ska-Gładysz, 2003), 11α-hydroxy-1-dehydrotes-
tosterone (14) and 11α-hydroxyandrost-1,4-diene-
3,17-dione (15) (Ahmed et al., 1996).
Apart from the hydroxylation, two other redox
reactions took place in the transformations pro-
moted by B. bassiana. The metabolites with satu-
rated C4,C5 or C1,C4 bonds were identified in tes-
tosterone (1) and 1-dehydrotestosterone (4)
transformations, respectively. The configuration of
5βfor compound 9was determined from the nega-
tive Cotton effect observed in the CD spectrum
([Θ]
304
=Ð1450). The presence of C3 and C17
carbonyl groups in the metabolite 11 resulted a
total positive Cotton effect, therefore the 5βcon-
figuration was confirmed by optical rotation mea-
surement ([α]
25
589
= + 72.9), which was in a good
agreement with the literature data (Allard, 1965).
Similar reduction of the C4,C5 double bound to the
5βconfiguration was observed for Beauveria globu-
lifera (Protiva et al., 1968).
Conversion of the alcohol at C17 into the ketones
1and 4also occurred. The presence of C17αmethyl
group inhibited reduction of both C4,C5 double bond
in 17α-methyltestosterone (2) and C1,C2 double
bond in 1-dehydro-17α-methyltestosterone (5).
It is noteworthy that the products of C4,C5
double bond reduction and/or C17 oxidation were
not found by transformation of 19-nortestosterone
(3). This substrate was relatively poorly metabo-
lized, which is in agreement with the results ob-
tained by Shibahara et al. (1970). They observed
that in spite of the induction of Aspergillus ochra-
ceus hydrolase only a low level of 11α-hydroxylation
of 19-nortestosterone could be achieved.
We have found that the main profile of biotrans-
formation of progesterone (6)byB. bassiana is the
E. Huszcza et al. · Steroid Transformations by Beauveria bassiana 107
side chain cleavage. Thus, the metabolite of proges-
terone was found to be the derivative of testoster-
one 8. Interestingly, 11α-hydroxytestosterone (8),
which was formed as the single product in high
yield, was not observed for other B. bassiana strains.
The main result of our study is the fact that struc-
tural differences in the 4-ene-3-oxo-steroids sub-
strates not effect the regio- and stereoselectivity of
the hydroxylation process. The steroid skeleton was
always attacked only at α-face of C11. Although the
6βand 11αpositions are expected to be equivalent
in enzyme-substrate complexes, 6β-hydroxy and
6β,11α-dihydroxy products were not found in any
our experiments.
The correlation between the structure of the sub-
stituent at C17 and the site specifity of hydroxyla-
tion of different steroid compounds by known 11α-
hydroxylators e.g. Rhizopus nigricans (Z
ˇakelj-Nar-
vic
ˇand Belic
ˇ, 1987), Aspergillus ochraceus (Tan and
Smith, 1968) and Cephalosporium aphidicola
(Boynton et al., 1997) was previously investigated.
Unlike in our study, it was shown that the side chain
at C17 had a strong influence on the position and
yield of hydroxylation by these fungi.
Introduction of the C5,C6 double bond to the ste-
roid skeleton slightly alters the transformation
course by the B. bassiana culture. Apart from the
major metabolite, 5-androstene-3β,11α,17β-triol
(17), incubation of dehydroepiandrosterone (7) with
B. bassiana gave a small quantity of 7α-hydroxyde-
hydroepiandrosterone (18). 7α-Hydroxylation was
confirmed by a down-field shift of the 6-H signal
(0.28 ppm as compared to substrate) and the exis-
tence of a narrow signal of 7β-H at 3.95 ppm. The
presence of the other minor product, androstenediol
(19), suggests that the 11α-hydroxylation was fol-
lowed by the reduction of the 17-ketone to the 17β-
alcohol. The hydroxylation of dehydroepiandroster-
one (7) at position 11αwithout further oxidation at
C3 is a typical feature of 11α-hydroxylating fungi
such as Rhizopus nigricans (Raspe and Richler,
1960), Rhizopus arrhizus (Holland and Diakow,
1979) and Aspergillus niger (Bell et al., 1972).
7α-Hydroxydehydroepiandrosterone (18) and an-
drostenediol (19) are metabolites of dehydroepian-
drosterone (7), which were detected in several mice
tissues. Compound 18 was described as a more po-
tent activator of immune processes in mice than 7
(Morfin and Courchay, 1994).
To sum up, we have found Beauveria bassiana to
be an efficient 11α-hydroxylator of dehydroepian-
drosterone (7) and 4-ene-3-ones, especially of 17α-
methyltestosterone (2) and 1-dehydro-17α-methyl-
testosterone (5). Compared to many other microor-
ganisms, Beauveria bassiana showed particularly
high regioselectivity and very low substrate specifity
in steroid hydroxylation. This phenomenon was ob-
served also for B. bassiana catalysed hydroxylations
of a variety of substrates e.g. amides, lactams, carba-
mates, azides and sulfonamides (Grogan and Hol-
land, 2000). Additionally, we have identified a new
dehydroepiandrosterone (7) transformation pro-
duct: 5-androstene-3β,11α,17β-triol (17).
108 E. Huszcza et al. · Steroid Transformations by Beauveria bassiana
Ahmed F., Williams A. D., and Smith K. E. (1996), Mi- Huszcza E. and Dmochowska-Gładysz J. (2003), Trans-
crobial transformations of steroids ÐX. Cytochromes formations of testosterone and related steroids in
P-450 11α-hydroxylase and C17-C20 lyase and A 1-ene Absidia glauca culture. J. Basic. Microbiol. 43, 113Ð
dehydrogenase transform steroids in Nectria haemato- 120.
cocca. J. Steroid Biochem. Mol. Biol. 58, 337Ð349. Jones E. R. H. (1973), The microbiological hydroxyla-
Allard S. (ed.) (1965), Tables of Constants and Numeri- tion of steroids and related compounds. Pure Appl.
cal Data. 14. Selected Constants, Optical Rotary Chem. 32,39Ð52.
Power, Ia. Steroids. Pergamon Press, Oxford, UK. Kirk D. N., Toms H. C., Douglas C., White K. A., Smith
Bayunova V. I., Gabinskaya K. N., Kolyanova T. S., Ko- K. E., Latif S., and Hubbard R. W. (1990), A survey
robova Y. N., and Grinenko G. S. (1989), Beauveria of the high-field
1
H NMR spectra of the steroids
sp. II. Transformation of androst-4-ene-3,17-dione and hormones, their hydroxylated derivatives, and re-
androst-1,4-diene-3,17-dione. Khim. Farm. Zh. 23, lated compounds. J. Chem. Soc. Perkin Trans. 2,
471Ð473. 1567Ð1594.
Bell A. M., Browne J. W., Denny W. A., Jones E. R. H., Krueger S. R. and Roberts D. W. (1997), Soil treatment
Kasal A., and Meakins G. D. (1972), Microbial hy- with entomopathogenic fungi for corn rootworm
droxylation of steroids. Part VI. Hydroxylation of sim- (Diabrotica spp.) larval control. Biol. Control 9,67Ð
ple mono- and di-oxygenated 5α-androstanes and of 74.
3-oxoestranes with the fungus Aspergillus ochraceus. Morfin R. and Courchay G. (1994), Pregnenolone and
J. Chem. Soc. Perkin Trans. I 23, 2930Ð2936. dehydroepiandrosterone as precursors of native 7-hy-
Bidochka M. J., Kasperski J. E., and Wild G. A. M. droxylated metabolites which increase the immune re-
(1998), Occurrence of the entomopathogenic fungi sponse in mice. J. Steroid Biochem. Mol. Biol. 50,
Metarhizium anisopliae and Beauveria bassiana in 91Ð100.
soils from temperate and near-northern habitats. Can. Mulock B. and Chandler L. (2000), Field-cage studies of
J. Bot. 76, 1198Ð1204. Beauveria bassiana (Hyphomycetes: Moniliaceae) for
Bing L. A. and Lewis L. C. (1991), Suppression of Os- the suppression of adult western corn rootworm, Dia-
tinia nubilalis (Huebner) (Lepidoptera: Pyralidae) by brotica vigifera vigifera (Cleoptera: Chrysomelidae).
endophytic Beauveria bassiana (Balsamo) Vuillemin. Biocontrol Sci. Technol. 10,51Ð60.
Environ. Entomol. 20, 1207Ð1211. Protiva J., Schwarz V., and Martinkova J. (1968), Forma-
Bing L. A. and Lewis L. C. (1992), Endophytic Beau- tion of 5β-saturated steroid derivatives by the 11α-
veria bassiana (Balsamo) Vuillemin in corn: the influ- hydroxylating fungus Beauveria globulifera. Folia
ence of the plant growth stage and Ostinia nubilalis Microbiol. 13, 139Ð145.
(Huebner). Biocontrol. Sci. Technol. 2,39Ð47. Raspe G. and Richler H. (1960), Ger. Patent
Boynton J., Hanson J. R., and Hunter A. C. (1997), The No. 1,080,553.
hydroxylation of some 13α-methylsteroids by Cepha- Sanada V., Fajkos J., and Protiva J. (1977), Microbial ox-
losporium aphidicola. Phytochemistry 45, 951Ð956. ygenation of norsteroids with Beauveria bassiana.
C
ˇapek A. and Fassatiova O. (1977), Some chemical char- Collect. Czech. Chem. Commun. 42, 3646Ð3653.
acteristic of species of the genus Beauveria. Folia Schubert A., Heller K., Onken D., Schwarz S., Siebert
Microbiol. (Prague) 22, 308Ð310. R., and Zaumseil J. (1962), Conversion of steroids
C
ˇapek A., Hanc O., Tadra M., and Tuma J. (1966), Mi- with microorganisms. Z. Chem. 2, 289Ð297.
crobial transformation of steroids XXV. Effect of sub- Shibahara M., Moody J. A., and Smith L. L. (1970),
stituents on the direction of the transformation of ste- Microbial hydroxylations V. 11α-Hydroxylation of
roids by Beauveria bassiana. Folia Microbiol. (Prague) progesterone by cell-free preparations of Aspergillus
11, 159Ð162. ochraceus. Biochim. Biophys. Acta 202, 172Ð179.
Griffiths D. A., Brown D. E., and Jezequel S. G. (1993), Smith K. E., Latif S., and Kirk D. N. (1990), Microbial
Metabolism of xenobiotics by Beauveria bassiana. transformations of steroids ÐVI. Transformation of
Xenobiotica 23, 1085Ð1100. testosterone and androstenedione by Botryosphaerica
Grogan G. J. and Holland H. L. (2000), The biocatalytic obtusa. J. Steroid Biochem. 35, 115Ð120.
reactions of Beauveria spp. J. Mol. Cat 9,1Ð32. Tan L., and Smith L. (1968), Microbial hydroxylations
Hajek A. E. and St. Leger R. J. (1994), Interactions be- II. Oxidative cleavage of the 17β-acetyl side chain of
tween fungal pathogens and insect hosts. Annu. Rev. progesterone by Aspergillus ochraceus. Biochim. Bio-
Entomol. 39, 293Ð322. phys. Acta 152, 758Ð769.
Holland H. L. and Diakow P. R. P. (1979), Microbial hy- Z
ˇakelj-Narvic
ˇM. and Belic
ˇI. (1987), Hydroxylation of
droxylation of steroids. 5. Metabolism of androst-5- steroids with 11α-hydroxylase of Rhizopus nigricans.
ene-3,17-dione and related compounds by Rhizopus J. Steroid Biochem. 28, 197Ð201.
arrhizus ATCC 11145. Can. J. Chem. 57, 436Ð440.
... Beauveria bassiana (i) 5-Androsten-3β,11α,17β-triol (ii) 7α-Hydroxy dehydroepiandrosterone [23] 1-Dehydro-17αmethyltestosterone (7) Beauveria bassiana 11α-Hydroxy-1-dehydro-17αmethyltestosterone ...
... Macrophomina phaseolina (i) 17β-Hydroxy-17αmethyl-5α-andros-1-en-3,11dione (ii) 14α,17β-Dihydroxy-17αmethyl-5α-androstan-3,11dione (iii) 17β-Hydroxy-17αmethyl-5α-andros-1,14-dien-3,11-dione (iv) 17β-Hydroxy-17αmethyl-5α-androstan-3,11-dione (v) 11α-Hydroxymestanolone (11α,17β-dihydroxy-17αmethyl-5α-androstan-3-one) Mesterolone (12) Cunninghamella blakesleeana (i) 1α-Methyl-1β,11β,17βtrihydroxy-5α-androstan-3-one (ii) 1α-Methyl-7α,11β,17βtrihydroxy-5α-androstan-3-one (iii) 1α-Methyl-1β,6α,17βtrihydroxy-5α-androstan-3-one (iv) 1α-Methyl-1β,11α,17βtrihydroxy-5α-androstan-3-one (v) 1α-methyl-11α,17β-dihydroxy-5α-androstan-3-one (vi) 1α-methyl-6α,17β-dihydroxy-5α-androstan-3-one (vii) 1α-methyl-7α,17βdihydroxy-5α-androstan-3-one [26] Macrophomina phaseolina 1α-Methyl,17β-hydroxy-5α-androstan-3,6-dione [26] Cephalosporium aphidicola (i) (1α, 5α)-1-Methylandrostane-3,17-dione (ii) (1α, 5α, 15α)-15-Hydroxy-1methylandrostane-3,17-dione [27] Fusarium lini (i) (5α)-1-Methylandrost-1-en-3,17-dione (ii) (1α, 5α, 6α, 17β)-6,17-Dihydroxy-1-methylandrostan-3-one (iii) (1α, 5α, 15α, 17β)-15,17-Dihydroxy-1-methylandrostan-3-one (iv) (5α, 15α, 17β) Cunninghamella elegans (i) 6β,17β-Dihydroxy-17αmethylandrost-1,4-dien-3-one (ii) 15α,17β-Dihydroxy-17αmethylandrost-1,4-dien-3-one (iii) 11α,17β-Dihydroxy-17αmethylandrost-1,4-dien-3-one (iv) 6β,12β,17β-Trihydroxy-17αmethylandrost-1,4-dien-3-one (v) 6β,15α,17β-Trihydroxy-17αmethylandrost-1,4-dien-3-one [28] Macrophomina phaseolina (i) 17β-Hydroxy-17αmethylandrost-1,4-dien-3,6dione (ii) 7β,17β-Dihydroxy-17αmethylandrost-1,4-dien-3-one (iii) 15β,17β-Dihydroxy-17αmethylandrost-1,4-dien-3-one (iv) 17β-Hydroxy-17αmethylandrost-1,4-dien-3,11dione (v) 11β,17β-Dihydroxy-17αmethylandrost-1,4-dien-3-one Beauveria bassiana 11α -Hydroxy-19-nortestosterone [23] Cunninghamella echinulata (i) 10β,12β,17β-trihydroxy-19nor-4-androsten-3-one (ii) 10β,16α,17β-trihydroxy-19nor-4-androsten-3-one (iii) 6β,10β,17β-trihydroxy-19nor-4-androsten-3-one (iv) 10β,17β-dihydroxy-19-nor-4androsten-3-one (v) 6β,17β-dihydroxy-19-nor-4androsten-3-one [30] Cunninghamella blakesleeana (i) 6β,10β,17β-trihydroxy-19nor-4-androsten-3-one (ii) 10β,17β-dihydroxy-19-nor-4androsten-3-one (iii) 10β-hydroxy-19-nor-4androsten-3,17-dione (iv) 16β,17β-dihydroxy-19-nor-4androsten-3-one [30] Substrate Microorganism Product % Yield * Reference Norandrostenedione (19) Fusarium culmorum (i) 6β-Hydroxy-19-nortestosterone (ii) 6β-Hydroxy-19norandrostenedione [16] Corynespora melonis 9α-Hydroxy-19-norandrostenedione [31] Nocardia restrictus 9α-Hydroxy-19-norandrostenedione [31] Oxandrolone (20) Rhizopus stolonifer ...
... Oxymetholone (21) Macrophomina phaseolina (i) 17β-Hydroxy-2-(hydroxymethyl)-17α-methyl-5α-androstan-1-en-3-one (ii) 2α,17α-Di(hydroxymethyl)-5αandrostan-3β-17β-diol [33] Rhizopus stolonifer 2α,17α-Di(hydroxymethyl)-5αandrostan-3β-17β-diol [33] Fusarium lini (i) 17β-Hydroxy-2-(hydroxymethyl)-17α-methyl-5α-androstan-1-en-3-one (ii) 17α-Methyl-5α-androstan-2α,3β-17β-triol (iii) 17β-Hydroxy-2-(hydroxymethyl)-17αmethylandrost-1,4-dien-3-one [33] Testosterone (22) Beauveria bassiana (i) 11α-Hydroxytestosterone (ii) 5α-Androstan-11α,17β-diol-3one (iii) 11α-Hydroxyandrost-4-ene-3,17-dione (iv) 5α-Androstan-11α-ol-3,17dione [23] Substrate Microorganism Product % Yield * Reference Rhizopus stolonifer (i) Androst-4-en-3,17-dione (ii) Testolactone (iii) 17-Hydroxy-5α-androstan-1,6-dione (iv) 11α-Hydroxyandrost-4-en-3,17-dione (v) 11α-Hydroxytestolactone [34] Fusarium lini ...
Article
Microbial biotransformations of various anabolic steroids are reviewed. Studies on oxidation, reduction, and carbon bond cleavage are highlighted. Various anabolic steroid substrates, their metabolites and the microorganisms used for the biotransformations are compiled covering the literature from the period 1984−2018.
... The possibility of using B. bassiana against a wide spectrum of insects has made it the object of numerous studies, and over time also its catalytic properties have been investigated. Many studies have shown that the B. bassiana enzyme apparatus has the ability to transform a variety of substrates including aromatic amines [17], amino acids [18], terpenes [19], flavonoids [20] and steroids [21][22][23] and is able to carry out, among other reactions, hydroxylation [21,22], acetylation [17], epoxidation [24], Baeyer-Villiger oxidation [22,23], glycosylation [17,25], sulfoxidation [18], dealkylation, reduction and ester hydrolysis [26]. Among the biotransformations of steroid compounds carried out in B. bassiana cultures, the following transformations have been described: hydroxylation at the 11α [23,26] or 7α position [21] and Baeyer-Villiger D-ring oxidation [22,23]. ...
... The possibility of using B. bassiana against a wide spectrum of insects has made it the object of numerous studies, and over time also its catalytic properties have been investigated. Many studies have shown that the B. bassiana enzyme apparatus has the ability to transform a variety of substrates including aromatic amines [17], amino acids [18], terpenes [19], flavonoids [20] and steroids [21][22][23] and is able to carry out, among other reactions, hydroxylation [21,22], acetylation [17], epoxidation [24], Baeyer-Villiger oxidation [22,23], glycosylation [17,25], sulfoxidation [18], dealkylation, reduction and ester hydrolysis [26]. Among the biotransformations of steroid compounds carried out in B. bassiana cultures, the following transformations have been described: hydroxylation at the 11α [23,26] or 7α position [21] and Baeyer-Villiger D-ring oxidation [22,23]. ...
... Many studies have shown that the B. bassiana enzyme apparatus has the ability to transform a variety of substrates including aromatic amines [17], amino acids [18], terpenes [19], flavonoids [20] and steroids [21][22][23] and is able to carry out, among other reactions, hydroxylation [21,22], acetylation [17], epoxidation [24], Baeyer-Villiger oxidation [22,23], glycosylation [17,25], sulfoxidation [18], dealkylation, reduction and ester hydrolysis [26]. Among the biotransformations of steroid compounds carried out in B. bassiana cultures, the following transformations have been described: hydroxylation at the 11α [23,26] or 7α position [21] and Baeyer-Villiger D-ring oxidation [22,23]. B. bassiana KCh BBT and B. bassiana KCh J1 strains were used for biotransformation of DHEA (dehydroepiandrosterone). ...
Article
Full-text available
Background: Steroid compounds with a 6,19-oxirane bridge possess interesting biological activities including anticonvulsant and analgesic properties, bacteriostatic activity against Gram-positive bacteria and selective anti-glucocorticoid action, while lacking mineralocorticoid and progestagen activity. Results: The study aimed to obtain new derivatives of 3β-acetyloxy-5α-chloro-6,19-oxidoandrostan-17-one by microbial transformation. Twelve filamentous fungal strains were used as catalysts, including entomopathogenic strains with specific activity in the transformation of steroid compounds. All selected strains were characterised by high biotransformation capacity for steroid compounds. However, high substrate conversions were obtained in the cultures of 8 strains: Beauveria bassiana KCh BBT, Beauveria caledonica KCh J3.4, Penicillium commune KCh W7, Penicillium chrysogenum KCh S4, Mucor hiemalis KCh W2, Fusarium acuminatum KCh S1, Trichoderma atroviride KCh TRW and Isaria farinosa KCh KW1.1. Based on gas chromatography (GC) and nuclear magnetic resonance (NMR) analyses, it was found that almost all strains hydrolysed the ester bond of the acetyl group. The strain M. hiemalis KCh W2 reduced the carbonyl group additionally. From the P. commune KCh W7 and P. chrysogenum KCh S4 strain cultures a product of D-ring Baeyer-Villiger oxidation was isolated, whereas from the culture of B. bassiana KCh BBT a product of hydroxylation at the 11α position and oxidation of the D ring was obtained. Three 11α-hydroxy derivatives were obtained in the culture of I. farinosa KCh KW1.1: 3β,11α-dihydroxy-5α-chloro-6,19-oxidoandrostan-17-one, 3β,11α,19-trihydroxy-5α-chloro-6,19-oxidoandrostan-17-one and 3β,11α-dihydroxy-5α-chloro-6,19-oxidoandrostan-17,19-dione. They are a result of consecutive reactions of hydrolysis of the acetyl group at C-3, 11α- hydroxylation, then hydroxylation at C-19 and its further oxidation to lactone. Conclusions: As a result of the biotransformations, seven steroid derivatives, not previously described in the literature, were obtained: 3β-hydroxy-5α-chloro-6,19-oxidoandrostan-17-one, 3β,17α-dihydroxy-5α-chloro-6,19-oxidoandrostane, 3β-hydroxy-5α-chloro-17α-oxa-D-homo-6,19-oxidoandrostan-17-one, 3β,11α-dihydroxy-5α-chloro-17α-oxa-D-homo-6,19-oxidoandrostan-17-one and the three above-mentioned 11α-hydroxy derivatives. This study will allow a better understanding and characterisation of the catalytic abilities of individual microorganisms, which is crucial for more accurate planning of experiments and achieving more predictable results.
... As one of the most common biocatalysts, entomopathogenic fungus Beauveria bassiana has multiple applications, especially in the hydroxylation of saturated and aromatic carbon atoms, Baeyer-Villiger/sulfide oxidation, keto-alcohol/alkene redox reaction, heteroatom dealkylation and epoxide, ester hydrolysis and glucosidation [8][9][10][11][12]. In our earlier study, a wild-type B. bassiana ZJB16001 [7, 13] capable of hydroxylating (R)-2phenoxypropionic acid (POPA) at C-4 site into HPOPA, was obtained from environmental samples and the derivative strain B. bassiana CCN-A7 was obtained after multi-round mutagenesis. ...
... In many organisms, reactive oxygen species (ROS) are highly reactive molecules mainly including superoxide radicals (O 2− ), hydrogen peroxide (H 2 O 2 ), and hydroxyl radicals (HO − ), all of which function in regulating various physiological processes [12,[17][18][19]. There has been increasing evidence that oxidative stress could cause the posttranslational modification of enzymes and thus change enzymes activity rapidly and reversibly. ...
Article
Full-text available
This study is the first report on production and characterization of the enzyme from an Ornithinibacillus species. A 4.2-fold increase in the extracellular protease (called L9T) production from Ornithinibacillus caprae L9T was achieved through one factor at-a-time approach and response surface methodological optimization. L9T protease exhibited a unique protein band with a mass of 25.9 kDa upon sodium dodecyl sulfate-polyacrylamide gel electrophoresis. This novel protease was active over a range of pH (4-13), temperatures (30-80 oC) and salt concentrations (0-220 g/L), with the maximal activity observed at pH 7, 70 oC and 20 g/L NaCl. Proteolytic activity was upgraded in the presence of Ag+, Ca2+ and Sr2+, but was totally suppressed by 5 mM phenylmethylsulfonyl fluoride which suggests that this enzyme belongs to the serine protease family. L9T protease was resistant to certain common organic solvents and surfactants; particularly, 5 mM Tween 20 and Tween 80 improved the activity by 63 and 15%, respectively. More importantly, L9T protease was found to be effective in dehairing of goatskins, cowhides and rabbit-skins without damaging the collagen fibers. These properties confirm the feasibility of L9T protease in industrial applications, especially in leather processing.
... When wild-type microbial cells were used, the microbial 17βreduction of 17-oxosteroids was usually accompanied with other reactions such as hydroxylation, reduction of double bond and dehydrogenation [9][10][11][12][13].Through microorganism screening, a strain Zygowilliopsis sp. WY7905 was found to convert 17-oxosteroids to 17β-hydroxysteroids without concurrence of other reactions. ...
... This may be due to the steric effect of ethyl group on C13 (Table 3, entries 9 and 10). The enzymes showed low activity with steroids having a C3-hydroxyl group (Table 3, entries [11][12][13], suggesting that the functional group on steroidal A-ring also affects the reactivity of wtRasADH and mutant enzymes. Compared to other substrates, the catalytic efficiency of F205I and F205A toward 5 and 8 showed more significant changes relative to wtRasADH, 80.6-fold and 91.8-fold activity enhancement was observed for F205I, and 175-fold and 66.4-fold improvement for F205A. ...
Article
Enzymatic stereospecific reduction of 17-oxosteroids offers an attractive approach to access 17β-hydroxysteroids of pharmaceutical importance. In this study, by adjusting the flexibility of α6-helix at the substrate entrance of the alcohol dehydrogenase from Ralstonia sp. (RasADH), the catalytic activity toward the stereospecific 17β-reduction of androstenedione was improved without sacrifice of the enantioselectivity. Among the mutants, F205I and F205A exhibited up to 623- and 523-fold improvement in catalytic efficiency, respectively, towards a range of different 17-oxosteroids compared to the wild-type enzyme. The corresponding 17β-hydroxysteroids were prepared in optically pure form with high space-time productivity and isolated yields using F205I as the biocatalyst, indicating that these mutants are promising biocatalysts for this useful transformation. These results suggest that modulating the flexibility of the active site lid offers an effective approach to engineer alcohol dehydrogenase for accommodating bulky steroidal substrates.
... As one of the most common biocatalysts, entomopathogenic fungus Beauveria bassiana has multiple applications, especially in the hydroxylation of saturated and aromatic carbon atoms, Baeyer-Villiger/sulfide oxidation, keto-alcohol/alkene redox reaction, heteroatom dealkylation and epoxide, ester hydrolysis and glucosidation [8][9][10][11][12]. In our earlier study, a wild-type B. bassiana ZJB16001 [7, 13] capable of hydroxylating (R)-2phenoxypropionic acid (POPA) at C-4 site into HPOPA, was obtained from environmental samples and the derivative strain B. bassiana CCN-A7 was obtained after multi-round mutagenesis. ...
... In many organisms, reactive oxygen species (ROS) are highly reactive molecules mainly including superoxide radicals (O 2− ), hydrogen peroxide (H 2 O 2 ), and hydroxyl radicals (HO − ), all of which function in regulating various physiological processes [12,[17][18][19]. There has been increasing evidence that oxidative stress could cause the posttranslational modification of enzymes and thus change enzymes activity rapidly and reversibly. ...
Article
Ethanol often accumulates during the process of wine fermentation, and mitophagy hascritical role in ethanol output. However, the relationship between mitophagy and ethanolstress is still unclear. In this study, the expression of ATG11 and ATG32 genes exposed toethanol stress was accessed by real-time quantitative reverse transcription polymerase chainreaction (qRT-PCR). The result indicated that ethanol stress induced expression of the ATG11and ATG32 genes. The colony sizes and the alcohol yield of atg11 and atg32 were alsosmaller and lower than those of wild type strain under ethanol whereas the mortality ofmutants is higher. Furthermore, compared with wild type, the membrane integrity and themitochondrial membrane potential of atg11 and atg32 exhibited greater damage followingethanol stress. In addition, a greater proportion of mutant cells were arrested at the G1/G0 cellcycle. There was more aggregation of peroxide hydrogen (H2O2) and superoxide anion (O2•-)in mutants. These changes in H2O2 and O2•- in yeasts were altered by reductants or inhibitors of scavenging enzyme by means of regulating the expression of ATG11 and ATG32 genes. Inhibitors of the mitochondrial electron transport chain (mtETC) also increased production of H2O2 and O2•- by enhancing expression of the ATG11 and ATG32 genes. Further results showed that activator or inhibitor of autophagy also activated or inhibited mitophagy byaltering production of H2O2 and O2•. Therefore, ethanol stress induces mitophagy which improves yeast the tolerance to ethanol and the level of mitophagy during ethanol stress is regulated by ROS derived from mtETC.
Article
Chemicals and poisons in the body interfere with the cell cycle and inhibit the growth of cancer cells. In this way, the function of chemicals in the body is controlled by taking anti-cancer drugs. Due to the degradability and compatibility of carbon nanotubes and boron nitride with the environment, they can act as suitable drug carriers for the transfer of anticancer drugs and deliver the drugs to the target cells. In the current work, the encapsulation of Formestane (FMS) anticancer drug into the carbon (CNT) and boron nitride (BNNT) (8,8) nanotubes was investigated for the first time using the density functional theory: B3LYP/3-21G* and the natural bond orbital analysis in the gas phase. Using natural bond orbital analysis, the charge transfer between FMS drug and CNT and BNNT nanotubes (8,8)/ FMS (BNNT/FMS) complexes were explored. Based on the results obtained from the calculation of encapsulation energy, it was found that the adsorption process was favorable. The interaction effects of FMS drug and CNT and BNNT (8,8) nanotubes on the natural bond orbital charge, the chemical shift parameters, and electronic properties were also evaluated. This study revealed that CNT and BNNT (8,8) nanotubes can be a suitable carrier for FMS drug delivery. The ultraviolet-visible spectra of the FMS drug, the CNT and BNNT (8,8), and the BNNT/FMS complexes were computed using time-dependent density functional theory (DFT: B3LYP) calculations.
Article
Full-text available
(R)-2-(4-hydroxyphenoxy)propionic acid (HPOPA) is a key intermediate for the preparation of aryloxyphenoxypropionic acid herbicides (R-isomer). In order to improve the HPOPA production from the substrate (R)-2-phenoxypropionic acid (POPA) with Beauveria bassiana CCN-A7, static cultivation and H2O2 addition were attempted and found to be conducive to HPOPA production. It was the first report on HPOPA production under static cultivation and reactive oxygen species (ROS) induction. In this premise, the cultivation conditions and fermentation medium compositions were optimized. As a result, the optimal carbon source, organic nitrogen source, and inorganic nitrogen source were determined to be glucose, peptone, and ammonium sulfate, respectively. The optimal inoculum size and fermentation temperature were 13.3% and 28°C, respectively. The significant factors including glucose, peptone, and H2O2, identified based on Plackett-Burman design, were further optimized through Central Composite Design (CCD). The optimal concentrations/amounts were as follows: glucose 38.81 g/L, peptone 7.28 g/L, and H2O2 1.08 mL/100 mL. Under the optimized conditions, HPOPA titer was improved from 9.6 g/L to 19.53 g/L, representing an increase of 2.03-fold. The results obtained in this work will provide novel strategies for improving the hydroxy aromatics biosynthesis.
Article
R-2-(4-hydroxyphenoxy)propionicacid (HPOPA) is a valuable intermediate for the synthesis of enantiomerically pure aryloxyphenoxypropionic acid herbicides. In this work, to improve the HPOPA biosynthesis by Beauveria bassiana ZJB16002 from the substrate R-2-phenoxypropionic acid (POPA), the original HPOPA producer B. bassiana ZJB16002 was subjected to physical mutagenesis with 137 Cs-γ irradiation and chemical mutagen N-methyl-N'-nitro-N-nitrasoguanidine (NTG) induced mutagenesis. The effects of different treatment doses of the mutagens on the lethal rate and positive mutation rate were investigated, and the results showed that the optimal 137 Cs-γirradiation dose and NTG concentration was 850 Gy and 500 μg/mL, respectively. Under these conditions, a mutant strain CCN-7 with the highest HPOPA production capacity was obtained through two rounds of 137 Cs-γ irradiation treatment followed by one round of NTG mutagenesis. At the substrate (POPA) concentration of 50 g/L, HPOPA titer of CCN-7 reached 36.88 g/L, which was 9.73-fold higher than the parental strain. The morphology of the wild-type and mutant strain was compared and the results might provide helpful information in exploration of the correlation of morphology and biochemical features of B. bassiana. This article is protected by copyright. All rights reserved.
Article
The occurrence of deuteromycetous entomopathogenic fungi was examined in 266 soil samples representing 86 locations across temperate and near northern habitats in Ontario, Canada. Entomopathogenic fungi were isolated by baiting the soil with waxworm larvae, Galleria mellonella L., and incubating at 8, 15, or 25 degrees C. Entomopathogenic fungi were isolated from 91% of the locations sampled across Ontario. The most abundant species were Metarhizium anisopliae (Metschn.) Sorok. (357 isolates) and Beauveria bassiana (Bals.) Vuill. (187 isolates). Thirteen isolates of Paecilomyces spp. were also found. Beauveria bassiana was isolated more frequently in soils from near northern locations, relative to M. anisopliae. Beauveria bassiana was isolated more frequently from larvae baited in soils incubated at 8 and 15 degrees C, while M. anisopliae was isolated most frequently at 25 degrees C. Thus, B. bassiana is more psychrophilic than M. anisopliae. From 47 of the locations in a temperate area (southern Ontario and the Kawartha Lakes region), two sites, one from an agricultural habitat and one from a natural habitat, were sampled within 1 km of each other. In these locations, B. bassiana was predominantly recovered more often from soils of natural habitats, while M. anisopliae was recovered more often in agricultural habitats. The occurrence of M. anisopliae and B. bassiana was not related to soil type or pH.
Article
Fungi of the Beauveria spp., typified by Beauveria bassiana ATCC 7159, are among the most frequently used whole cell biocatalysts. They have been reported to catalyse many different reactions, including oxidative, reductive and hydrolytic transformations, of a wide range of substrates. This review covers the range and application of biocatalytic reactions of Beauveria spp., with emphasis on the scope and utility of Beauveria-catalysed reactions for preparative biotransformations.
Article
The efficacy of the entomopathogenic fungus Beauveria bassiana (Balsamo) Vuillemin, was tested as a control agent for adult western corn rootworm, Diabrotica virgifera virgifera LeConte, in walk-in field cages. Suspensions of B. bassiana conidia were applied to corn plants in cages into which laboratory-reared beetles had been released. Beetles were collected at 3 and 5 days post-application and evaluated in the laboratory for mortality. Mortality was 10, 29 and 50%, at rate equivalents of 7 X 10 12 , 2 X 10 13 (two applications), and 5 X 10 13 conidia/ ha, respectively. There was no significant difference in mortality of beetles collected at 3 days compared with 5 days post-application. Mortality due to B. bassiana was 24% when beetles were released into field cages 24 h post-application (5 X 10 13 conidia/ha) compared with 50% when beetles were present during the application. Beetle mortality declined significantly with increasing time from application in feeding assays carried out with leaf samples removed from plants at 0, 12, 24 and 72 h post-application. Mortality of beetles collected from treated plants within cages and maintained in the laboratory was found to overestimate the population decline by 10% when compared with beetle estimates from treated plants within field cages.
Article
Three groups of 3-keto-4-ene steroids with different side chains were used as substrates for the induced 11α -hydroxylase of Rhizopus nigricans. The highest total bioconversion as well as the highest yield of 11α-hydroxylated product is found using progesterone as substrate. By changing the polarity of the side chain, much higher yields of 6β- and 7β-hydroxylated products relative to 11α-hydroxylated product are obtained. Our results thus provide evidence for the importance of the side chain in steroid-enzyme interactions.
Article
1 H NMR chemical shifts are presented for virtually all the protons in 166 steroids. These comprise mainly the hormones testosterone, androst-4-ene-3,17-dione, progesterone, and a wide range of their hydroxylated derivatives, some corticosteroids including aldosterone and a series of its derivatives, together with miscellaneous steroids comprising a variety of androstane and pregnane derivatives, bile acids, and sterols, to provide the first extensive collection of data for use in correlating 1H chemical shifts with structure.
Article
Field trials were conducted in 1988 and 1989 to determine the efficacy of soil applications of the entomopathogenic fungiMetarhizium anisopliaeandBeauveria bassianaagainst the southern corn rootworm (SCR),Diabrotica undecimpunctata,in corn. Dry mycelial particles were incorporated into the soil to a depth of 15 cm at rates of 9.3 and 0.93 g particles/m row at planting time. The concentration of fungus in soil increased during the first month postapplication and then decreased during the second month for both trials. One month after the application of fungus, the treated soil adjacent to corn plants was infested with 300–350 and 600–650 SCR eggs during 1988 and 1989, respectively. In 1988, the high treatment rate ofM. anisopliaewas equal to the uninfested control in preventing goosenecked plants and larval feeding on roots and in limiting adult emergence. The effectiveness of the high rate ofM. anisopliaecoincided with a stable soil concentration of the fungus of at least 2.7 × 105CFU/cm3soil during larval development. All other fungal treatments provided lesser degrees of plant protection and reduced adult emergence compared to the SCR-infested control. In 1989, the effectiveness of all fungal treatments was reduced compared to that in 1988. However, root damage ratings among plants in fungus-treated soil were significantly lower than those in the SCR-infested control.
Article
The products of the incubation of androst-5-ene-3,17-dione (2a), 3β-hydroxyandrost-5-ene-17-one (2b), and androsta-3,5-diene-17-one (3) with Rhizopus arrhizus ATCC 11145 under a variety of conditions have been identified and the mechanisms of their formation discussed. In addition, several C-(4,5)- and C-(5,6)-epoxyandrostanes have been incubated with R. arrhizus, the products identified and possible pathways for their formation presented.
Article
The occurrence of deuteromycetous entomopathogenic fungi was examined in 266 soil samples representing 86 locations across temperate and near northern habitats in Ontario, Canada. Entomopathogenic fungi were isolated by baiting the soil with waxworm larvae, Galleria mellonella L., and incubating at 8, 15, or 25°C. Entomopathogenic fungi were isolated from 91% of the locations sampled across Ontario. The most abundant species were Metarhizium anisopliae (Metschn.) Sorok. (357 isolates) and Beauveria bassiana (Bals.) Vuill. (187 isolates). Thirteen isolates of Paecilomyces spp. were also found. Beauveria bassiana was isolated more frequently in soils from near northern locations, relative to M. anisopliae. Beauveria bassiana was isolated more frequently from larvae baited in soils incubated at 8 and 15°C, while M. anisopliae was isolated most frequently at 25°C. Thus, B. bassiana is more psychrophilic than M. anisopliae. From 47 of the locations in a temperate area (southern Ontario and the Kawartha Lakes region), two sites, one from an agricultural habitat and one from a natural habitat, were sampled within 1 km of each other. In these locations, B. bassiana was predominantly recovered more often from soils of natural habitats, while M. anisopliae was recovered more often in agricultural habitats. The occurrence of M. anisopliae and B. bassiana was not related to soil type or pH.Key words: Metarhizium, Beauveria, entomopathogenic fungi, fungal population biology, soil ecology.