ArticlePDF Available

Crystal structures of free and ligand‐bound forms of the TetR/AcrR‐like regulator SCO3201 from Streptomyces coelicolor suggest a novel allosteric mechanism

Wiley
The FEBS Journal
Authors:

Abstract and Figures

TetR/AcrR‐like transcription regulators enable bacteria to sense a wide variety of chemical compounds and to dynamically adapt the expression levels of specific genes in response to changing growth conditions. Here, we describe the structural characterisation of SCO3201, an atypical TetR/AcrR family member from Streptomyces coelicolor that strongly represses antibiotic production and morphological development under conditions of overexpression. We present crystal structures of SCO3201 in its ligand‐free state as well as in complex with an unknown inducer, potentially a polyamine. In the ligand‐free state, the DNA‐binding domains of the SCO3201 dimer are held together in an unusually compact conformation and, as a result, the regulator cannot span the distance between the two half‐sites of its operator. Interaction with the ligand coincides with a major structural rearrangement and partial conversion of the so‐called hinge helix (α4) to a 310‐conformation, markedly increasing the distance between the DNA‐binding domains. In sharp contrast to what was observed for other TetR/AcrR‐like regulators, the increased interdomain distance might facilitate rather than abrogate interaction of the dimer with the operator. Such a ‘reverse’ induction mechanism could expand the regulatory repertoire of the TetR/AcrR family and may explain the dramatic impact of SCO3201 overexpression on the ability of S. coelicolor to generate antibiotics and sporulate.
This content is subject to copyright. Terms and conditions apply.
Crystal structures of free and ligand-bound forms of the
TetR/AcrR-like regulator SCO3201 from Streptomyces
coelicolor suggest a novel allosteric mechanism
Sebastiaan Werten
1
, Paul Waack
2,
*, Gottfried J. Palm
2
, Marie-Jo
elle Virolle
3
and
Winfried Hinrichs
2
1 Institute of Genetic Epidemiology, Medical University of Innsbruck, Austria
2 Institute of Biochemistry, University of Greifswald, Germany
3 Institute for Integrative Biology of the Cell (I2BC), Universit
e Paris-Saclay, CEA, CNRS, Gif-sur-Yvette, France
Keywords
gene regulation; polyamine; SAXS;
transcription; X-ray crystallography
Correspondence
S. Werten, Institute of Genetic
Epidemiology, Medical University of
Innsbruck, Sch
opfstr. 41, 6020 Innsbruck,
Austria
Tel: +43 512900370578
E-mail: sebastiaan.werten@i-med.ac.atW.
Hinrichs, Institute of Biochemistry,
University of Greifswald, Felix-Hausdorff-Str.
4, 17489 Greifswald, Germany
Tel: +49 17654172520
E-mail: winfried.hinrichs@uni-greifswald.de
Present address
*Bayer AG, Berlin, Germany
(Received 14 May 2022, revised 14 July
2022, accepted 25 August 2022)
doi:10.1111/febs.16606
TetR/AcrR-like transcription regulators enable bacteria to sense a wide
variety of chemical compounds and to dynamically adapt the expression
levels of specific genes in response to changing growth conditions. Here, we
describe the structural characterisation of SCO3201, an atypical TetR/
AcrR family member from Streptomyces coelicolor that strongly represses
antibiotic production and morphological development under conditions of
overexpression. We present crystal structures of SCO3201 in its ligand-free
state as well as in complex with an unknown inducer, potentially a polya-
mine. In the ligand-free state, the DNA-binding domains of the SCO3201
dimer are held together in an unusually compact conformation and, as a
result, the regulator cannot span the distance between the two half-sites of
its operator. Interaction with the ligand coincides with a major structural
rearrangement and partial conversion of the so-called hinge helix (a4) to a
3
10
-conformation, markedly increasing the distance between the DNA-
binding domains. In sharp contrast to what was observed for other TetR/
AcrR-like regulators, the increased interdomain distance might facilitate
rather than abrogate interaction of the dimer with the operator. Such a ‘re-
verse’ induction mechanism could expand the regulatory repertoire of the
TetR/AcrR family and may explain the dramatic impact of SCO3201 over-
expression on the ability of S. coelicolor to generate antibiotics and sporu-
late.
Introduction
Streptomyces are a genus of gram-positive, soil-
dwelling bacteria, capable of producing a wide variety
of secondary metabolites of immediate biomedical
interest. These include an estimated 80% of currently
available antibiotics, as well as numerous antitumour
agents, antifungals and antivirals [13]. Certain species
of Streptomyces have in addition gained popularity as
hosts for heterologous protein expression [4]. A char-
acteristic feature of streptomycetes is their complex
differentiation cycle, which involves profound changes
in metabolism as well as in morphology [57]. The dif-
ferentiation process is triggered by variations in the
intra- and extracellular concentrations of specific com-
pounds, particularly nutrients that constitute vital
sources of phosphorus and nitrogen [8]. Concentra-
tions of these substances are often gauged by one-
component systems [9], such as the TetR/AcrR [10]
family of transcription regulators. Members of this
Abbreviations
HTH, helix-turn-helix motif; SAXS, small-angle X-ray scattering; RMSD, root-mean-square deviation.
521The FEBS Journal 290 (2023) 521–532 Ó2022 The Authors. The FEBS Journal published by John Wiley & Sons Ltd on behalf of
Federation of European Biochemical Societies.
This is an open access article under the terms of the Creative Commons Attribution-NonCommercial-NoDerivs License, which permits use and
distribution in any medium, provided the original work is properly cited, the use is non-commercial and no modifications or adaptations are made.
family act as homodimeric repressors or activators,
which bind to palindromic sequences in the promoters
of target genes and are released by a cognate small
molecule, the inducer. TetR/AcrR-like regulators are
bipartite and consist of an N-terminal DNA-binding
domain that comprises a helix-turn-helix (HTH) motif,
followed by a C-terminal domain required for dimeri-
sation and inducer recognition [11]. In all cases where
crystal structures of DNA-bound as well as inducer-
bound forms are available, interaction with the small
molecule is found to allosterically increase the distance
between the DNA-binding domains, which then
exceeds the spacing between the half-sites in the opera-
tor [10,12]. As a consequence, DNA-binding is abro-
gated and transcription enabled [13].
SCO3201, a TetR/AcrR-like regulator from Strepto-
myces coelicolor, was originally described by Xu et al.
[14], who screened a genomic library for factors that
influence antibiotic production and development.
Overexpression of SCO3201 was found to strongly
reduce levels of actinorhodin (ACT), a polyketide
antibiotic, as well as to inhibit morphological differen-
tiation and sporulation of S. coelicolor [14]. The pro-
tein was shown to physically interact with the
promoters of at least 16 genes, including some that
encode downstream regulators, thereby directly and
indirectly affecting numerous genes that play a role in
the differentiation process [15]. An initial crystallo-
graphic analysis of SCO3201 at 2.1
A resolution
revealed a highly asymmetric homodimer, with an
unidentified small molecule bound to one of its sub-
units [15]. In this study, we describe a new crystal
form that lacks the unknown ligand, leading to an
essentially symmetric and unusually compact confor-
mation of the dimer. This compact conformation is
also observed in solution by means of small-angle X-
ray scattering (SAXS). Furthermore, we have repeated
the crystal structure determination of the asymmetric,
ligand-bound dimer at a higher resolution (1.89
A),
providing more detailed insight into the nature of the
small molecule, a possible polyamine. Taken together,
our data suggest that a novel allosteric mechanism
underlies the ligand response of SCO3201.
Results
The crystal structure of the ligand-free form of
SCO3201 reveals an unusually compact ‘closed’
state
The asymmetric unit of the ligand-free P2
1
crystal
form that we obtained contains four polypeptide
chains (AD), which form two homodimers (AB and
CD). The overall fold of SCO3201 (Fig. 1) is typical
of the TetR/AcrR family of transcription factors [10]
and comprises an all-helical regulatory domain that
also mediates dimerisation, preceded by a DNA-
binding three-helix-bundle (a1a3). The second and
third helix of this DNA-binding domain correspond to
a classical helix-turn-helix (HTH) motif. A distinguish-
ing feature of SCO3201 is the additional a-helix (a8)
within a solvent-exposed loop of the canonical TetR/
AcrR fold, which juts out from the core of the ligand-
binding domain (Fig. 1A). However, the most striking
aspect of the structure is arguably the exceptionally
short distance separating the HTH motifs of the
homodimer. The recognition helices (a3), which are
crucial for sequence-specific DNA binding [10], are
22.6
A apart in dimer AB and 24.5
A in dimer CD
(this is measured as the distance between the centres of
mass of the helical residues, amino acid residues 66
72, considering main-chain atoms only). In most struc-
tures of ligand-free and DNA-bound TetR/AcrR-like
regulators the corresponding distance is within the
range 3541
A, while even longer distances have been
observed for inducer-bound repressors [16]. The posi-
tioning of the DNA-binding domains of SCO3201,
which gives the dimer a remarkably compact (‘closed’)
appearance compared to other members of the TetR/
AcrR family, mainly arises because of a pronounced
bend in the so-called hinge helix (a4). This helix plays
a key role in the allosteric induction mechanism of
TetR/AcrR-like repressors, as it physically links the
HTH motif to the ligand-binding domain. Interest-
ingly, the equivalent of a4 in SCO3201 is a composite
helix, as its N-terminal part in direct contact with the
DNA-binding domain adopts a 3
10
-conformation. This
results in a bend of approximately 30°at the junction
with the a-helical section, in good agreement with the
average bend angle for 3
10
-atransitions (37°19°)
observed by Pal et al. [17]. Each of the four indepen-
dent protein chains in the crystal features a similarly
bent composite hinge helix and the AB and CD dimers
can be superimposed with an RMSD of 1.6
A for a
total of 358 aligned C
a
-atoms (Fig. 1B).
Ligand binding coincides with major
conformational changes
A second crystal form, belonging to space group
P2
1
2
1
2
1
, turned out to be indistinguishable from the
one that we had already analysed in earlier work [15].
However, crystals diffracted to higher resolution this
time (1.86
A versus 2.1
A). As in the earlier struc-
ture, the asymmetric unit comprises a single homod-
imer, whose subunits adopt markedly different
522 The FEBS Journal 290 (2023) 521–532 Ó2022 The Authors. The FEBS Journal published by John Wiley & Sons Ltd on behalf of
Federation of European Biochemical Societies.
The induction mechanism of SCO3201 S. Werten et al.
conformations. While chain B is virtually identical to
the four molecules in the ligand-free structure (with C
a
-
RMSD values ranging between 0.79 and 1.0
A), the
DNA-binding domain of chain A has moved in the out-
ward direction by approximately 10
A, resulting in an
increased a3a3 distance of 34.8
A (Fig. 2). This
change coincides with the binding of an unknown
ligand to a cavity within the A-subunit (Figs 2and 3).
The movement of the DNA-binding domain is caused
by a repositioning of the composite hinge helix, which
undergoes a 10°outward rotation (away from the
ligand binding pocket) as well as a 3
A shift in the
upward direction in Fig. 2A. This movement is accom-
panied by a rearrangement of the adjacent helices a5,
a6 and a7. All of these changes result from direct con-
tacts between the helices and the ligand, which force the
binding pocket to expand. Interestingly, the upward
shift of the hinge helix would cause its outer 3
10
-
segment to collide with the adjacent a7 helix, while the
DNA-binding domain would be drawn into the ligand-
binding domain. Such clashes are avoided by an exten-
sion of the 3
10
-helix (Fig. 2B), which now includes two
additional residues belonging to the a-helical region in
the ligand-free structure (Ala81 and Pro82). The result
is not only an appreciable elongation of the hinge helix,
but also a reduction of the bend angle to approximately
15°. This straightening is a direct consequence of the
fact that the remodelled 3
10
-ajunction does not obey
the sequence requirements for composite helices that
were formulated by Pal et al. [17]. In particular, Pro82
no longer occupies the favourable second position in
the a-helix, but instead acts as the final residue of the
3
10
-helix. This precludes proper capping of the a-helix
and formation of a larger bend. Instead, Pro82 is rigidly
integrated into the structural framework of the 3
10
-helix
by hydrogen bonds between its side chain (C
d
-H) and
the carbonyl moieties of Glu78 and Ala79 (with C
d
-O
distances of 3.1 and 3.4
A, respectively).
In contrast to the extensive changes to helices
a4a7, the main dimerisation interface consisting of
helices a9 and a10 remains essentially unaffected by
the presence of the ligand. These helices can be super-
imposed onto the corresponding part of the ligand-free
AB dimer with a C
a
-RMSD of 0.38
A (68 aligned resi-
dues). In the ligand-free dimer, however, two highly
similar salt bridges related by non-crystallographic
symmetry (Fig. 3B) connect the carboxylate of Glu155
in helix a7 to the guanidinium moiety of Arg195 in
helix a9 of the dimerisation partner. In the liganded
structure, the head group of the unknown small mole-
cule is at 2.5
A, that is well within hydrogen bonding
distance, from one of the carboxylate oxygen atoms of
Glu155. This interaction abrogates one of the hydro-
gen bonds between Glu155 and the side chain of
Arg1950, thereby breaking the electrostatic symmetry
of the two salt bridges.
Fig. 1. The structure of the ligand-free
SCO3201 homodimer. (A) Ribbon
representation of the dimer in two
orientations. Chains A and B are shown in
beige and green, respectively. The exposed
helix a8, which has no equivalent in other
TetR/AcrR regulators, is highlighted in
brown (chain A) and dark green (chain B).
Designations LBD and DBD indicate the
ligand-binding and DNA-binding domains,
respectively. The distance separating the
recognition helices (a3) of the two HTH-
motifs is indicated by an arrow. (B) Stereo
view of the superposed AB (colours as in
panel A) and CD (grey) homodimers, shown
as C
a
-traces.
523The FEBS Journal 290 (2023) 521–532 Ó2022 The Authors. The FEBS Journal published by John Wiley & Sons Ltd on behalf of
Federation of European Biochemical Societies.
S. Werten et al. The induction mechanism of SCO3201
The elusive ligand may be an aliphatic polyamine
Unfortunately, mass spectrometry analysis of dissolved
crystals failed to identify the unknown ligand. However,
owing to the markedly improved resolution of the data,
the contours of the small molecule are now clearly
defined in the electron density map (Fig. 3). Although
the identity of the ligand cannot be unambiguously
deduced from the density alone, the properties of the
binding site and the way it interacts with the ligand
necessitate a linear aliphatic chain of at least 8 heavy
(non-hydrogen) atoms, including a positively charged
(or at least polar) head group. In view of these require-
ments, aliphatic polyamines [18] seem to be likely candi-
dates. Well-balanced levels of these ubiquitous
compounds, the most abundant of which are sper-
midine, spermine and putrescine [18,19], are considered
essential for cell growth, proliferation and survival in
all organisms [20].InStreptomyces, polyamines have
been shown to play key roles in gene regulation and
metabolism [21]. On the basis of these considerations,
we have tentatively modelled spermidine into the elec-
tron density map. The observed Van der Waals contacts
to hydrophobic residues lining the binding pocket and
the short-distance interaction of the head group with
the carboxylate moiety of the Glu155 side chain are
consistent with this proposition. However, other
polyamines (including the closely related compounds
putrescine, cadaverine and spermine) are equally plausi-
ble candidates and further biochemical studies will be
required to unambiguously identify the natural ligand.
As the binding site widens towards the protein surface,
the outer region of the bound molecule may well be dis-
ordered and, as a result, remain invisible in the density
map. Consequently, the unknown molecule could be
considerably longer than the part that is actually
observed.
SCO3201 adopts the closed conformation in
solution
The markedly different conformations observed in
our crystal structures prompted us to characterise
SCO3201 in solution by means of SAXS (Fig. 4).
Comparison of the extrapolated forward scattering
(I
0
) to that of a bovine serum albumin (BSA) refer-
ence sample suggests a particle mass of 46.4 kDa,
while the concentration-independent Bayesian method
of Hajizadeh et al. [22], yields a value of 46.7 kDa.
These results are in excellent agreement with the pre-
dicted mass of the SCO3201 homodimer (46.7 kDa).
A more detailed analysis of the data using CRYSOL
[23] (Fig. 4A) indicates that the scattering from
SCO3201 is explained almost to within experimental
error (v
2
=1.2) by the CD dimer from the ligand-free
crystal structure. In contrast, using the asymmetric
liganded structure as a model (either with or without
the small molecule) leads to a considerable lack of fit
(v
2
=2.6). The same is true for a symmetrical ‘doubly
liganded’ model, constructed by superimposing a copy
of the A chain of the liganded structure onto the
dimerisation interface of the B chain and combining
that with the original A chain (v
2
=2.5). An ab initio
model calculated from the experimental data using
Fig. 2. Comparison of the ligand-bound and ligand-free SCO3201 structures. (A) Superposition of liganded (protein chains shown as beige
and green ribbons with the unknown small molecule, tentatively modelled as spermidine, represented by red spheres) and free SCO3201
(dimer AB, grey ribbons). The outward movement of one of the HTH motifs by 10
A, which coincides with ligand binding to the same pro-
tein chain, is indicated by an arrow. The portion of the hinge helix that undergoes substantial conformational changes (bright green and blue
in the respective structures) is shown in more detail in panel (B). (B) Conformational changes within the composite hinge helix upon ligand
binding. The N-terminal 3
10
-segment is extended by two residues previously part of the a-helical region (Ala81 and Pro82), leading to stretch-
ing and straightening of the hinge helix. All side-chain atoms beyond C
b
have been removed to improve clarity, except those of Pro82.
Hydrogen bonds are shown as dashed lines.
524 The FEBS Journal 290 (2023) 521–532 Ó2022 The Authors. The FEBS Journal published by John Wiley & Sons Ltd on behalf of
Federation of European Biochemical Societies.
The induction mechanism of SCO3201 S. Werten et al.
the program DAMMIN [24] closely matches the dimer
from 5EFY (Fig. 4B). We therefore conclude that, in
our SAXS experiments, SCO3201 predominantly
adopts a closed conformation similar to the one seen
in the ligand-free crystal structure. Several things may
explain why the ligand-induced conformation is not
observed here. In the first place, the basic conditions
(pH 8.5) might favour dissociation of the complex, as
the pH of the solution could have a considerable
impact on the protonation state of both the ligand
and the glutamate residue it is in contact with.
Indeed, crystals of the complex were obtained at pH
5.0, whereas the ligand-free protein crystallised at pH
8.5. Compared to the crystallisation assays, the SAXS
experiments were also carried out at much lower pro-
tein concentrations, which would be expected to shift
the binding equilibrium towards the dissociated state.
Furthermore, only a small fraction of the purified
protein may have been liganded in the first place.
While this might not prevent crystallisation of the
liganded species, the unliganded form would in this
case dominate the SAXS signal. Finally, we cannot
exclude the possibility that the unknown ligand was
not co-purified from the expression host, but con-
tained in the chemicals used during crystal structure
determination, in particular the paraffin oil (a com-
plex and poorly defined mixture of organic com-
pounds) that we used for cryoprotection. However,
mass spectrometry analysis of a protein solution
extensively equilibrated with paraffin oil also failed to
identify the unknown ligand.
Discussion
The crystallographic structure of the ligand-free form
of SCO3201 reported in the present study reveals a
domain arrangement that, to our knowledge, has not
been encountered in other members of the TetR/AcrR
family. The DNA-binding domains of the protein
dimer are unusually close to one another, mainly as a
result of a pronounced inward bend in the hinge
helices. We have several reasons to believe that this
arrangement is not merely the result of crystal packing
effects, but corresponds to the native state of the
ligand-free regulator. First, a transition from a 3
10
-
helix to an a-helix is fully expected to produce the
Fig. 3. The interaction of SCO3201 with the
inducer. (A) Stereo image of the liganded
cavity. The view is from the opening of the
cavity at the protein surface towards the
interior of the dimer, with the dyad axis
oriented vertically in the plane of the image.
Difference density (mFo-DFc) corresponding
to the inducer, contoured at 3.0 r, is shown
in blue. Nearby side chains are represented
as stick models, as is the spermidine
molecule that was modelled into the
density. The short-distance interaction
between the head group of the inducer and
the side chain of Glu155 is represented by
a dashed line. (B) The dimerisation interface
of ligand-bound SCO3201, viewed along the
dyad axis. Hydrogen bonds involving the
inducer, Glu155 and Arg195 are shown as
dashed lines, together with the
corresponding interatomic distances. As can
be seen here, the local non-crystallographic
symmetry of the dimerisation interface (the
Glu155-Arg1950and Glu1550-Arg195 salt
bridges in particular) is broken by the
interaction of Glu155 with the ligand.
525The FEBS Journal 290 (2023) 521–532 Ó2022 The Authors. The FEBS Journal published by John Wiley & Sons Ltd on behalf of
Federation of European Biochemical Societies.
S. Werten et al. The induction mechanism of SCO3201
kind of kinked junction that is observed here [17]. Sec-
ond, all four of the independent polypeptide chains in
the crystal feature a composite hinge helix with a com-
parable bend, despite the fact that each chain experi-
ences a different lattice environment. And finally, our
SAXS data corroborate the existence of the closed
conformation in solution.
Intriguingly, the compact arrangement of the two
HTH-motifs, with their recognition helices at less than
25
A from each other, would preclude binding to con-
secutive instances of the major groove on one face of
the promoter. This DNA-binding mode, which is a
hallmark of the TetR/AcrR family [10], requires an
a3a3 distance very close to 35
A[16]. This implies
that the SCO3201 dimer has to undergo a major con-
formational change if it is to interact with the opera-
tor. Although DNA binding by members of the TetR/
AcrR family is typically accompanied by appreciable
structural changes in the regulatory domain [10], all
currently known cases involve DNA-binding domains
that are far apart in the free repressor and need to be
brought into closer proximity to enable promoter
binding. Inducers can then act simply by preventing
such rearrangements [25]. Strikingly, this is the exact
opposite of what we seem to observe for SCO3201.
In our second crystal form, interaction with a small-
molecule ligand causes a large (~10
A) increase in the
distance between the DNA-binding domains, which
would now be compatible with DNA binding. This
observation suggests that SCO3201 could possibly func-
tion by means of a ‘reverse’ induction mechanism, in
the sense that ligand binding might increase, rather than
decrease, the affinity of the regulator for DNA. To our
knowledge, such a mechanism has not yet been sug-
gested for any naturally occurring TetR/AcrR family
members. Mutagenesis studies of TetR(D), however,
have given rise to artificial variants of this repressor (re-
ferred to as ‘revTetR’) that, instead of being released
from DNA by tetracycline like the wild-type protein,
show enhanced promoter binding in the presence of the
antibiotic [2628]. For the related LacI/GalR family of
transcription factors, cases of increased as well as
decreased DNA-binding in the presence of cognate
small-molecule ligands have been reported [29].
It is currently unclear whether or not the SCO3201
dimer is able to interact with a second inducer mole-
cule, which might further enlarge the physical distance
between the HTH motifs and lead to a complex tran-
scriptional response to increasing ligand concentra-
tions. It should be noted, however, that many TetR/
AcrR-like repressors only bind a single ligand molecule
per homodimer [10], either in a central cavity around
the symmetry axis of the dimer, as in FrrA [25],orin
one of several available pockets within the individual
subunits, for example in QacR [30] and TtgR [31].
The conformational changes that we observe in
SCO3201 are ultimately caused by a slight expansion
of the binding pocket, a direct consequence of the
bulkiness of the inducer. This effect should be virtually
independent of the precise binding specificity of the
cavity, a phenomenon that we have also observed for
the induction mechanisms of TetR(D) [32] and FrrA
[25]. As noted before, modularity of ligand recognition
and allostery is expected to result in a considerable
degree of evolutionary flexibility, allowing existing
transcriptional regulators to mutate and adapt to new
inducers and regulatory circuits while their induction
mechanism remains intact [32].
Loop-to-helix transitions within helix a4, sometimes
linked to the presence of helix-destabilising residues
(Pro, Gly) and generally triggered by direct interac-
tions of the helix with the inducer, have been reported
for several TetR/AcrR-like repressors [33,34]. The
allosteric mechanism of SCO3201, in contrast, relies
on a composite hinge helix and the elongation of its
Fig. 4. SAXS analysis of SCO3201 in solution. (A) Experimental
SAXS data (black dots) and theoretical curves calculated for the
ligand-free (red) and ligand-bound (blue) crystallographic structures.
Theoretical curves were fitted to the experimental data using CRY-
SOL, with 7PT0 and the CD dimer of 5EFY as the models, respec-
tively. (B) Ab initio model (grey surface) with the superposed 5EFY
CD dimer shown in ribbon representation (beige and green).
526 The FEBS Journal 290 (2023) 521–532 Ó2022 The Authors. The FEBS Journal published by John Wiley & Sons Ltd on behalf of
Federation of European Biochemical Societies.
The induction mechanism of SCO3201 S. Werten et al.
3
10
-segment in response to the expansion of the ligand-
binding cavity. Interconversion of 3
10
- and a-helices is
believed to be an important regulatory mechanism in
biomolecular signalling, in particular in voltage-gated
channels and other membrane-associated switches [35].
However, we are not aware of such mechanisms hav-
ing been described for transcription factors. A search
of the PDB using DALI [36] indicates that SCO3201
most strongly resembles entry 2RAE (Z-score: 16.4),
which corresponds to the 2.2
A crystal structure of a
repressor from Rhodococcus sp. RHA1. Although this
protein lacks the additional a-helix within loop a7-a8,
the relative positioning of most of its helices is roughly
comparable to that of counterparts in SCO3201
(Fig. 5A). Like SCO3201, 2RAE possesses a composite
hinge helix, with a helix-destabilising residue (Gly67)
at the transition between the a- and 3
10
-segments
(Fig. 5B). As in SCO3201, the 3
10
-segment also con-
tains a proline residue (Pro65), whose side chain con-
tributes to the 3
10
hydrogen-bonding network.
Although no ligands are present in the deposited
model, inspection of the 2mFo-DFc map reveals an
area of poorly explained electron density within the
putative binding cavity, presumably arising from the
presence of a planar organic molecule (Fig. 5B). This
observation strongly suggests that 2RAE corresponds
to the induced form of the repressor. Unlike the
ligand-bound form of SCO3201, both subunits of the
2RAE homodimer are liganded (they are related by
crystallographic symmetry), leading to an a3-a3 dis-
tance that appears to be too large for DNA binding
(41.8
A). Despite this intriguing difference, SCO3201
and 2RAE together constitute a novel category of reg-
ulators whose allosteric mechanisms exploit the special
properties of 3
10
-ajunctions in composite helices [17].
A more detailed analysis of these molecular switches
awaits the structural characterisation of complexes of
the proteins with promoter DNA, as well as the unam-
biguous identification of their natural inducers.
Involvement of SCO3201 in polyamine sensing, pos-
sibly by means of a reverse induction mechanism, may
explain the dramatic effect of overexpression of the
regulator on morphological differentiation and antibi-
otic production in S. coelicolor [14]. A recent study
has demonstrated that Streptomyces are able to detox-
ify and assimilate polyamines very efficiently, enabling
growth even under conditions where these molecules
constitute the sole nitrogen source [37]. Moreover,
polyamines are naturally abundant in the environ-
ments where Streptomyces are typically found [21].In
S. coelicolor,de novo biosynthesis of putrescine and
spermidine has been reported in the late-stationary
phase in rich medium, while cadaverine synthesis
occurred under conditions of iron limitation [38].
These observations strongly suggest that polyamines
play key roles in cell homeostasis, with depletion
likely to act as a nutritional stress signal. As
soon as the polyamine concentration falls below a
Fig. 5. Comparison of SCO3201 to 2RAE, the most similar PDB entry identified by DALI [36]. (A) Superposition of 2RAE (chains related by
crystallographic symmetry in dark blue and purple) onto the ligand-free SCO3201 dimer (chains C and D, grey). The hinge helices (a4) of
2RAE have been highlighted in light blue and pink. The kink in the hinge helix of 2RAE, around residue Gly67, is indicated by arrows. (B)
Detailed view of the composite hinge helix of 2RAE (pink stick model, side-chain atoms beyond C
b
removed except in Pro65) with the rest
of the protein chain in ribbon representation (purple). Hydrogen bonds are shown as dashed lines. Electron density (2mFo-DFc) likely to cor-
respond to an unidentified ligand, contoured at 1.2 r, is shown in blue. The authors of 2RAE interpreted this density as four water mole-
cules at less than 2.2
A from each other (red spheres).
527The FEBS Journal 290 (2023) 521–532 Ó2022 The Authors. The FEBS Journal published by John Wiley & Sons Ltd on behalf of
Federation of European Biochemical Societies.
S. Werten et al. The induction mechanism of SCO3201
certain threshold, genes involved in morphological
development and antibiotic production would need to
be derepressed, triggering differentiation and antibi-
otic production. Overexpression of SCO3201 would
amplify polyamine-dependent repression (and poten-
tially prolong it) as a result of the law of mass action.
Moreover, sequestering of polyamine by the overex-
pressed regulator might protect the former from being
metabolised, allowing maintenance of repression
throughout growth [14]. However, the fact that dele-
tion of SCO3201 has no detectable impact on morpho-
logical differentiation and antibiotic production [14]
suggests that control of the relevant genes involves
additional factors besides SCO3201. Two regulators of
polyamine catabolism have been described in S. coeli-
color: EpuRII (SCO5656), which controls the glnA3
gene encoding a c-glutamylpolyamine synthetase, and
the global regulator of nitrogen metabolism, GlnR,
which has been shown to interact with the glnA2 gene,
encoding a second c-glutamylpolyamine synthetase
[37]. Interestingly, the expression of glnR can only be
induced by cadaverine [37], whereas the expression of
EpuRII can be induced by putrescine, cadaverine and
spermidine [39]. It is therefore conceivable that several
dedicated factors sense different polyamines or groups
of polyamines in the context of an intricate regulatory
network that mediates cross-talk between a variety of
nutritional signals. Apart from sensors of nitrogen
sources [4043], such a network is likely to include fac-
tors that respond to the presence of inorganic phos-
phate [4446] and polyphosphates [4749].
Materials and methods
Protein expression and purification
In all experiments, a SCO3201 deletion mutant was used
that lacks the 20-residue region of low complexity at the
N-terminus of the full-length sequence, which we found
interfered with crystallisation. The construct, corresponding
to amino acids 21216 of NCBI reference sequence WP_
011028825.1, was expressed and purified essentially as
described earlier [15]. In brief, the His
6
-tagged protein was
overproduced in Escherichia coli strain C41 using a modi-
fied pET-28b vector that instead of the thrombin recogni-
tion sequence encodes a TEV-protease site. Cultures in LB
medium containing 50 lgmL
1
kanamycin were grown
shaking at 37 °C until they reached an OD (600 nm) of
0.8, followed by a further 4 h incubation in the presence of
1m
MIPTG. Upon centrifugation (15 min at 10 000 g,
4°C), cells from 1 L of culture were resuspended in 30 mL
lysis buffer (20 mMTris/HCl pH 7.0, 400 mMNaCl, 20 mM
of imidazole and 1 mMof DTT) and disrupted by means of
sonication for 8 min on ice using a Sonopuls HD 2070
(Bandelin Electronic GmbH, Berlin, Germany). Sonication
was repeated twice, with 2 min of cooling between cycles.
Cell debris was removed by centrifugation for 1 h at
40 000 gand the cleared lysate applied to a Ni
2+
-charged
Poros 20 MC column (Thermo Fisher Scientific GmbH,
Dreieich, Germany). The protein was eluted using lysis buf-
fer containing 250 mMof imidazole. TEV protease was
added directly to the eluate to a final concentration of
0.1 mgmL
1
. The solution was then incubated for 3 days
at 4 °C, leading to complete removal of the His
6
-tag.
SCO3201 was further purified by gel filtration (Superdex
75, GE Healthcare GmbH, Munich, Germany) in a buffer
containing 20 mMTris/HCl pH 8.5, 400 mMNaCl, 50 mM
imidazole and 1 mMDTT. The protein was concentrated to
30 mgmL
1
using a Vivaspin 15 concentrator (Sartorius
Lab Instruments GmbH, G
ottingen, Germany) with a
10 kDa molecular weight cut-off. The concentrate was
flash-frozen in 50 lL of aliquots using liquid nitrogen and
stored at 80 °C until further use. Total protein yield was
approximately 100 mg from 1 L of culture. The protein
appeared as a single band on Coomassie-stained SDS/
PAGE gels.
X-ray crystallography
Crystals of ligand-free SCO3201 belonging to space group
P2
1
were obtained using the hanging drop vapour diffusion
method at 20 °C, with a reservoir solution containing 12%
ethanol and 0.1 MTris/HCl pH 8.5. For cryoprotection,
crystals were covered in paraffin oil (alkane chain
length 20, Carl Roth GmbH, Karlsruhe, Germany) prior
to flash freezing in liquid nitrogen. Diffraction data to a
resolution of 2.7
A were collected at beamline BL14.2,
operated by the Joint Berlin MX-Laboratory at the BESSY
II electron storage ring (Berlin-Adlershof, Germany) [50].
Crystals of the liganded form of SCO3201, which
belonged to space group P2
1
2
1
2
1
, were also obtained via
hanging drop vapour diffusion at 20 °C. Ahead of crystalli-
sation assays, 1 mL protein solution (2 mgmL
1
) was
incubated with 10 lL of paraffin oil. The rationale behind
this approach was that the oil, which was used as a cry-
oprotectant in the determination of the earlier struc-
ture (4CGR), might contain the unknown ligand, in which
case pre-equilibration might increase occupancy and
improve crystal quality. The solution, without the organic
phase, was subsequently concentrated to 50 lL
(40 mgmL
1
) and used for crystallisation. The reservoir
solution in these experiments contained 1 Mof (NH
4
)
H
2
PO
4
and 0.1 Mof Tris/HCl pH 5.0. Again, paraffin oil
was used for cryoprotection. A dataset with a resolution of
1.86
A was collected at beamline P13, operated by EMBL
Hamburg at the PETRA III storage ring (DESY, Ham-
burg, Germany) [51].
528 The FEBS Journal 290 (2023) 521–532 Ó2022 The Authors. The FEBS Journal published by John Wiley & Sons Ltd on behalf of
Federation of European Biochemical Societies.
The induction mechanism of SCO3201 S. Werten et al.
All diffraction data were processed with XDS [52,53]. Both
crystal structures were solved using PHASER [54], in combina-
tion with the CCP4 suite [55]. For the ligand-free form, chain
B of protein data bank (PDB) entry 4CGR was used as the
molecular replacement search model. Searches with chain A
or the complete homodimer (AB) did not yield meaningful
solutions. For the liganded form, the dimer was used as the
search model. The structures were refined using the CCP4
suite [55] and REFMAC5[56,57]. All molecular graphics illus-
trations were produced with PYMOL 0.99rc6 (DeLano Scien-
tific LLC, Palo Alto, CA, USA). Centres of mass and
angles between helices were determined using CALCOM [58]
and QHELIX [59], respectively. Data collection and refinement
statistics can be found in Table 1.
SAXS
SAXS data were collected at beamline BM29 [60] of the
EMBL Outstation at ESRF, Grenoble, using a PILATUS
1Mpixel detector. Measurements covered the momentum
transfer range 0.0020.60
A
1
(s=4psin(h)/k, where 2his
the scattering angle and kis the X-ray wavelength,
0.992
A). All experiments took place at a temperature of
293 K. A capillary flow cell was used to avoid radiation
damage. Several SCO3201 dilutions (0.195.5 mgmL
1
)
were analysed in a buffer containing 20 mMTris/HCl pH
8.5, 400 mMNaCl, 50 mMimidazole and 1 mMDTT. All
data were normalised to the intensity of the transmitted
beam and radially averaged. Scattering of the buffer was
subtracted and the resulting curves were scaled to unity
protein concentration (1 mgmL
1
). For further data analy-
sis, version 3.0.3 of the ATSAS package was used [61]. Theo-
retical scattering patterns based on structural models were
fitted to the experimental data using the program CRYSOL,
which accounts for protein surface hydration [23].Ab initio
models were calculated using the program DAMMIN [24].
Acknowledgements
We would like to thank Martha E. Brennich (EMBL
Grenoble) for SAXS data collection.
Conflict of interest
The authors declare no conflict of interest.
Author contributions
SW, M-JV and WH conceived, initiated and super-
vised this study; PW expressed, purified and crys-
tallised SCO3201; PW and GJP collected X-ray
diffraction data; PW and WH solved and refined the
crystal structures; SW analysed the SAXS data; SW
interpreted the experimental results, analysed the
induction mechanism and produced the artwork; SW
and WH wrote the manuscript.
Peer review
The peer review history for this article is available at
https://publons.com/publon/10.1111/febs.16606.
Data availability statement
Atomic coordinates and X-ray diffraction data for the
crystal structures of ligand-free and induced SCO3201
have been deposited in the Protein Data Bank (PDB)
with accession codes 5EFY and 7PT0. SAXS data and
models have been deposited in the Small-Angle
Table 1. Data collection and structure refinement statistics. Values
in parentheses correspond to the highest resolution shell.
Protein or complex SCO3201
SCO3201/
spermidine
PDB ID 5EFY 7PT0
Data collection
Radiation source,
beamline
BESSY, BL14.2 PETRA III,
EMBL c/o
DESY, P13
Detector RAYONIX MX-225 Pilatus 6M
Wavelength (
A) 0.9184 0.9763
Temperature (K) 100 100
Resolution range (
A) 94.62.70 (2.852.70) 61.61.89
(2.001.89)
Unit cell parameters
a,b,c(
A)
a,b,c(°)
51.79, 98.43, 89.61
90, 100.78, 90
55.13, 79.88,
96.87
90, 90, 90
Space group P2
1
P2
1
2
1
2
1
Unique reflections 23 495 (3916) 34 934 (5288)
Redundancy 3.0 (2.9) 6.3 (5.5)
Mean I/r(I) 7.7 (1.4) 20.7 (1.1)
Completeness (%) 96.6 (97.7) 99 (94.1)
R
meas
(%) 6.1 (101.3) 4.0 (148)
Wilson B-factor (
A
2
) 79.1 59.5
CC 1/2 0.998 (0.279) 1 (0.53)
Refinement
R
cryst
/number of
reflections
0.258/22 236 0.204/33 204
R
free
/number of reflections 0.297/1227 0.220/1727
Number of
non-hydrogen
atoms, protein/
water/ligand
5309/23/3002/87/10
RMSD bond lengths (
A) 0.016 0.010
RMSD bond angles (°) 1.58 1.60
RMSD torsion
angles (°)
4.68 5.29
Ramachandran
parameters
(%), favoured/
allowed/outlier
99.5/0.5/96.6/2.8/0.6
Average B-factors (
A
2
) 77.6 64.6
529The FEBS Journal 290 (2023) 521–532 Ó2022 The Authors. The FEBS Journal published by John Wiley & Sons Ltd on behalf of
Federation of European Biochemical Societies.
S. Werten et al. The induction mechanism of SCO3201
Scattering Biological Data Bank (SASBDB) with
accession code SASDN75.
References
1 Proc
opio REL, da Silva IR, Martins MK, de Azevedo
JL, de Ara
ujo JM. Antibiotics produced by
Streptomyces. Braz J Infect Dis. 2012;16:46671.
2 Antoraz S, Santamar
ıa RI, D
ıaz M, Sanz D, Rodr
ıguez
H. Toward a new focus in antibiotic and drug discovery
from the Streptomyces arsenal. Front Microbiol.
2015;6:461.
3 Kemung HM, Tan LTH, Khan TM, Chan KG,
Pusparajah P, Goh BH, et al. Streptomyces as a
prominent resource of future anti-MRSA drugs. Front
Microbiol. 2018;9:2221.
4 Berini F, Marinelli F, Binda E. Streptomycetes:
attractive hosts for recombinant protein production.
Front Microbiol. 2020;11:1958.
5 Claessen D, de Jong W, Dijkhuizen L, W
osten HAB.
Regulation of Streptomyces development: reach for the
sky! Trends Microbiol. 2006;14:3139.
6Fl
ardh K, Buttner MJ. Streptomyces morphogenetics:
dissecting differentiation in a filamentous bacterium.
Nat Rev Microbiol. 2009;7:3649.
7Fl
ardh K, Richards DM, Hempel AM, Howard M,
Buttner MJ. Regulation of apical growth and hyphal
branching in Streptomyces. Curr Opin Microbiol.
2012;15:73743.
8 McCormick JR, Fl
ardh K. Signals and regulators that
govern Streptomyces development. FEMS Microbiol
Rev. 2012;36:20631.
9 Romero-Rodr
ıguez A, Robledo-Casados I, S
anchez S.
An overview on transcriptional regulators in
Streptomyces. Biochim Biophys Acta BBA.
2015;1849:101739.
10 Cuthbertson L, Nodwell JR. The TetR family of
regulators. Microbiol Mol Biol Rev. 2013;77:44075.
11 Hinrichs W, Kisker C, D
uvel M, M
uller A, Tovar K,
Hillen W, et al. Structure of the Tet repressor-
tetracycline complex and regulation of antibiotic
resistance. Science. 1994;264:41820.
12 Orth P, Schnappinger D, Hillen W, Saenger W,
Hinrichs W. Structural basis of gene regulation by the
tetracycline inducible Tet repressor-operator system.
Nat Struct Biol. 2000;7:2159.
13 Aleksandrov A, Schuldt L, Hinrichs W, Simonson T.
Tet repressor induction by tetracycline: a molecular
dynamics, continuum electrostatics, and
crystallographic study. J Mol Biol. 2008;378:898912.
14 Xu D, Seghezzi N, Esnault C, Virolle M-J. Repression
of antibiotic production and sporulation in
Streptomyces coelicolor by overexpression of a TetR
family transcriptional regulator. Appl Environ
Microbiol. 2010;76:774153.
15 Xu D, Waack P, Zhang Q, Werten S, Hinrichs W,
Virolle MJ. Structure and regulatory targets of
SCO3201, a highly promiscuous TetR-like regulator of
Streptomyces coelicolor M145. Biochem Biophys Res
Commun. 2014;450:5138.
16 Haberl F, Lanig H, Clark T. Induction of the
tetracycline repressor: characterization by molecular-
dynamics simulations. Proteins. 2009;77:85766.
17 Pal L, Dasgupta B, Chakrabarti P. 3(10)-helix adjoining
alpha-helix and beta-strand: sequence and structural
features and their conservation. Biopolymers.
2005;78:14762.
18 Michael AJ. Polyamines in eukaryotes, bacteria, and
archaea. J Biol Chem. 2016;291:14896903.
19 Michael AJ. Polyamine function in archaea and
bacteria. J Biol Chem. 2018;293:18693701.
20 Wallace HM. The polyamines: past, present and future.
Essays Biochem. 2009;46:19.
21 Krysenko S, Matthews A, Busche T, Bera A,
Wohlleben W. Poly- and monoamine metabolism in
Streptomyces coelicolor: the new role of glutamine
synthetase-like enzymes in the survival under
environmental stress. Microb Physiol. 2021;31:23347.
22 Hajizadeh NR, Franke D, Jeffries CM, Svergun DI.
Consensus Bayesian assessment of protein molecular
mass from solution X-ray scattering data. Sci Rep.
2018;8:7204.
23 Svergun D, Barberato C, Koch MHJ. CRYSOLa
program to evaluate X-ray solution scattering of
biological macromolecules from atomic coordinates.
J Appl Cryst. 1995;28:76873.
24 Svergun DI. Restoring low resolution structure of
biological macromolecules from solution scattering
using simulated annealing. Biophys J. 1999;76:2879
86.
25 Werner N, Werten S, Hoppen J, Palm GJ, G
ottfert
M, Hinrichs W. The induction mechanism of the
flavonoid-responsive regulator FrrA. FEBS J.
2021;289:50718.
26 Kamionka A, Bogdanska-Urbaniak J, Scholz O, Hillen
W. Two mutations in the tetracycline repressor change
the inducer anhydrotetracycline to a corepressor.
Nucleic Acids Res. 2004;32:8427.
27 Scholz O, Henßler EM, Bail J, Schubert P, Bogdanska-
Urbaniak J, Sopp S, et al. Activity reversal of Tet
repressor caused by single amino acid exchanges. Mol
Microbiol. 2004;53:77789.
28 Resch M, Striegl H, Henssler EM, Sevvana M, Egerer-
Sieber C, Schiltz E, et al. A protein functional leap:
how a single mutation reverses the function of the
transcription regulator TetR. Nucleic Acids Res.
2008;36:4390401.
29 Swint-Kruse L, Matthews KS. Allostery in the LacI/
GalR family: variations on a theme. Curr Opin
Microbiol. 2009;12:12937.
530 The FEBS Journal 290 (2023) 521–532 Ó2022 The Authors. The FEBS Journal published by John Wiley & Sons Ltd on behalf of
Federation of European Biochemical Societies.
The induction mechanism of SCO3201 S. Werten et al.
30 Schumacher MA, Miller MC, Grkovic S, Brown MH,
Skurray RA, Brennan RG. Structural mechanisms of
QacR induction and multidrug recognition. Science.
2001;294:215863.
31 Alguel Y, Meng C, Ter
an W, Krell T, Ramos JL,
Gallegos MT, et al. Crystal structures of multidrug
binding protein TtgR in complex with antibiotics and
plant antimicrobials. J Mol Biol. 2007;369:82940.
32 Werten S, Schneider J, Palm GJ, Hinrichs W. Modular
organisation of inducer recognition and allostery in the
tetracycline repressor. FEBS J. 2016;283:210214.
33 Manjasetty BA, Halavaty AS, Luan CH, Osipiuk J,
Mulligan R, Kwon K, et al. Loop-to-helix transition in
the structure of multidrug regulator AcrR at the
entrance of the drug-binding cavity. J Struct Biol.
2016;194:1828.
34 Sawai H, Yamanaka M, Sugimoto H, Shiro Y, Aono S.
Structural basis for the transcriptional regulation of
heme homeostasis in Lactococcus lactis.J Biol Chem.
2012;287:3075568.
35 Vieira-Pires RS, Morais-Cabral JH. 310 helices in
channels and other membrane proteins. J Gen Physiol.
2010;136:58592.
36 Holm L. Using Dali for protein structure comparison.
Methods Mol Biol. 2020;2112:2942.
37 Krysenko S, Okoniewski N, Nentwich M, Matthews A,
B
auerle M, Zinser A, et al. A second gamma-
Glutamylpolyamine synthetase, GlnA2, is involved in
polyamine catabolism in Streptomyces coelicolor.Int J
Mol Sci. 2022;23:3752.
38 Burrell M, Hanfrey CC, Kinch LN, Elliott KA,
Michael AJ. Evolution of a novel lysine decarboxylase
in siderophore biosynthesis. Mol Microbiol.
2012;86:48599.
39 Krysenko S, Okoniewski N, Kulik A, Matthews A,
Grimpo J, Wohlleben W, et al. Gamma-
Glutamylpolyamine synthetase GlnA3 is involved in the
first step of polyamine degradation pathway in
Streptomyces coelicolor M145. Front Microbiol.
2017;8:726.
40 Feng W-H, Mao X-M, Liu Z-H, Li Y-Q. The ECF
sigma factor SigT regulates actinorhodin production in
response to nitrogen stress in Streptomyces coelicolor.
Appl Microbiol Biotechnol. 2011;92:100921.
41 Lewis RA, Shahi SK, Laing E, Bucca G, Efthimiou G,
Bushell M, et al. Genome-wide transcriptomic analysis
of the response to nitrogen limitation in Streptomyces
coelicolor A3(2). BMC Res Notes. 2011;4:78.
42 Tiffert Y, Franz-Wachtel M, Fladerer C, Nordheim A,
Reuther J, Wohlleben W, et al. Proteomic analysis of
the GlnR-mediated response to nitrogen limitation in
Streptomyces coelicolor M145. Appl Microbiol
Biotechnol. 2011;89:114959.
43 Zhu Y, Zhang P, Zhang J, Xu W, Wang X, Wu L,
et al. The developmental regulator MtrA binds GlnR
boxes and represses nitrogen metabolism genes in
Streptomyces coelicolor.Mol Microbiol. 2019;112:2946.
44 Mart
ın JF, Rodr
ıguez-Garc
ıa A, Liras P. The master
regulator PhoP coordinates phosphate and nitrogen
metabolism, respiration, cell differentiation and
antibiotic biosynthesis: comparison in Streptomyces
coelicolor and Streptomyces avermitilis.J Antibiot
(Tokyo). 2017;70:53441.
45 Yang R, Liu X, Wen Y, Song Y, Chen Z, Li J. The
PhoP transcription factor negatively regulates
avermectin biosynthesis in Streptomyces avermitilis.
Appl Microbiol Biotechnol. 2015;99:1054757.
46 Smirnov A, Esnault C, Prigent M, Holland IB, Virolle M-J.
Phosphate homeostasis in conditions of phosphate
proficiency and limitation in the wild type and the phoP
mutant of Streptomyces lividans.PLoS ONE. 2015;10:
e0126221.
47 Chouayekh H, Virolle M-J. The polyphosphate kinase
plays a negative role in the control of antibiotic
production in Streptomyces lividans.Mol Microbiol.
2002;43:91930.
48 Werten S, Rustmeier NH, Gemmer M, Virolle M-J,
Hinrichs W. Structural and biochemical analysis of a
phosin from Streptomyces chartreusis reveals a
combined polyphosphate- and metal-binding fold.
FEBS Lett. 2019;593:201929.
49 Shikura N, Darbon E, Esnault C, Deniset-Besseau A, Xu
D, Lejeune C, et al. The Phosin PptA plays a negative role
in the regulation of antibiotic production in Streptomyces
lividans.Antibiotics (Basel). 2021;10:325.
50 Mueller U, Darowski N, Fuchs MR, F
orster R,
Hellmig M, Paithankar KS, et al. Facilities for
macromolecular crystallography at the Helmholtz-
Zentrum Berlin. J Synchrotron Radiat. 2012;19:4429.
51 Cianci M, Bourenkov G, Pompidor G, Karpics I,
Kallio J, Bento I, et al. P13, the EMBL
macromolecular crystallography beamline at the low-
emittance PETRA III ring for high- and low-energy
phasing with variable beam focusing. J Synchrotron
Radiat. 2017;24:32332.
52 Kabsch W. XDS. Acta Crystallogr D Biol Crystallogr.
2010;66:12532.
53 Kabsch W. Integration, scaling, space-group assignment
and post-refinement. Acta Crystallogr D Biol
Crystallogr. 2010;66:13344.
54 McCoy AJ, Grosse-Kunstleve RW, Adams PD, Winn
MD, Storoni LC, Read RJ. Phaser crystallographic
software. J Appl Cryst. 2007;40:65874.
55 Winn MD, Ballard CC, Cowtan KD, Dodson EJ,
Emsley P, Evans PR, et al. Overview of the CCP4 suite
and current developments. Acta Crystallogr D Biol
Crystallogr. 2011;67:23542.
56 Winn MD, Murshudov GN, Papiz MZ. Macromolecular
TLS refinement in REFMAC at moderate resolutions.
Methods Enzymol. 2003;374:30021.
531The FEBS Journal 290 (2023) 521–532 Ó2022 The Authors. The FEBS Journal published by John Wiley & Sons Ltd on behalf of
Federation of European Biochemical Societies.
S. Werten et al. The induction mechanism of SCO3201
57 Murshudov GN, Vagin AA, Dodson EJ. Refinement of
macromolecular structures by the maximum-likelihood
method. Acta Crystallogr D Biol Crystallogr.
1997;53:24055.
58 Costantini S, Paladino A, Facchiano AM. CALCOM: a
software for calculating the center of mass of proteins.
Bioinformation. 2008;2:2712.
59 Lee HS, Choi J, Yoon S. QHELIX: a computational
tool for the improved measurement of inter-helical
angles in proteins. Protein J. 2007;26:55661.
60 Pernot P, Round A, Barrett R, de Maria Antolinos A,
Gobbo A, Gordon E, et al. Upgraded ESRF BM29
beamline for SAXS on macromolecules in solution. J
Synchrotron Radiat. 2013;20:6604.
61 Manalastas-Cantos K, Konarev PV, Hajizadeh NR,
Kikhney AG, Petoukhov MV, Molodenskiy DS, et al.
ATSAS 3.0: expanded functionality and new tools for
small-angle scattering data analysis. J Appl Cryst.
2021;54:34355.
532 The FEBS Journal 290 (2023) 521–532 Ó2022 The Authors. The FEBS Journal published by John Wiley & Sons Ltd on behalf of
Federation of European Biochemical Societies.
The induction mechanism of SCO3201 S. Werten et al.
Article
This article provides a comprehensive review and sequence‐structure analysis of transcription regulator (TR) families, TetR and OmpR / PhoB , involved in specialized secondary metabolite ( SSM ) biosynthesis and resistance. Transcription regulation is a fundamental process, playing a crucial role in orchestrating gene expression to confer a survival advantage in response to frequent environmental stress conditions. This process, coupled with signal sensing, enables bacteria to respond to a diverse range of intra and extracellular signals. Thus, major bacterial signaling systems use a receptor domain to sense chemical stimuli along with an output domain responsible for transcription regulation through DNA ‐binding. Sensory and output domains on a single polypeptide chain (one component system, OCS ) allow response to stimuli by allostery, that is, DNA ‐binding affinity modulation upon signal presence/absence. On the other hand, two component systems ( TCSs ) allow cross‐talk between the sensory and output domains as they are disjoint and transmit information by phosphorelay to mount a response. In both cases, however, TRs play a central role. Biosynthesis of SSMs , which includes antibiotics, is heavily regulated by TRs as it diverts the cell's resources towards the production of these expendable compounds, which also have clinical applications. These TRs have evolved to relay information across specific signals and target genes, thus providing a rich source of unique mechanisms to explore towards addressing the rapid escalation in antimicrobial resistance (AMR). Here we focus on the TetR and OmpR family TRs , which belong to OCS and TCS , respectively. These TR families are well‐known examples of regulators in secondary metabolism and are ubiquitous across different bacteria, as they also participate in a myriad of cellular processes apart from SSM biosynthesis and resistance. As a result, these families exhibit higher sequence divergence, which is also evident from our bioinformatic analysis of 158 389 and 77 437 sequences from TetR and OmpR family TRs , respectively. The analysis of both sequence and structure allowed us to identify novel motifs in addition to the known motifs responsible for TR function and its structural integrity. Understanding the diverse mechanisms employed by these TRs is essential for unraveling the biosynthesis of SSMs . This can also help exploit their regulatory role in biosynthesis for significant pharmaceutical, agricultural, and industrial applications.
Article
Full-text available
Streptomyces coelicolor is a soil bacterium living in a habitat with very changeable nutrient availability. This organism possesses a complex nitrogen metabolism and is able to utilize the polyamines putrescine, cadaverine, spermidine, and spermine and the monoamine ethanolamine. We demonstrated that GlnA2 (SCO2241) facilitates S. coelicolor to survive under high toxic polyamine concentrations. GlnA2 is a gamma-glutamylpolyamine synthetase, an enzyme catalyzing the first step in polyamine catabolism. The role of GlnA2 was confirmed in phenotypical studies with a glnA2 deletion mutant as well as in transcriptional and biochemical analyses. Among all GS-like enzymes in S. coelicolor, GlnA2 possesses the highest specificity towards short-chain polyamines (putrescine and cadaverine), while its functional homolog GlnA3 (SCO6962) prefers long-chain polyamines (spermidine and spermine) and GlnA4 (SCO1613) accepts only monoamines. The genome-wide RNAseq analysis in the presence of the polyamines putrescine, cadaverine, spermidine, or spermine revealed indication of the occurrence of different routes for polyamine catabolism in S. coelicolor involving GlnA2 and GlnA3. Furthermore, GlnA2 and GlnA3 are differently regulated. From our results, we can propose a complemented model of polyamine catabolism in S. coelicolor, which involves the gamma-glutamylation pathway as well as other alternative utilization pathways.
Article
Full-text available
Bradyrhizobium diazoefficiens, a bacterial symbiont of soybean and other leguminous plants, enters a nodulation‐promoting genetic programme in the presence of host‐produced flavonoids and related signalling compounds. Here, we describe the crystal structure of an isoflavonoid‐responsive regulator (FrrA) from Bradyrhizobium, as well as cocrystal structures with inducing and noninducing ligands (genistein and naringenin, respectively). The structures reveal a TetR‐like fold whose DNA‐binding domain is capable of adopting a range of orientations. A single molecule of either genistein or naringenin is asymmetrically bound in a central cavity of the FrrA homodimer, mainly via C–H contacts to the π‐system of the ligands. Strikingly, however, the interaction does not provoke any conformational changes in the repressor. Both the flexible positioning of the DNA‐binding domain and the absence of structural change upon ligand binding are corroborated by small‐angle X‐ray scattering (SAXS) experiments in solution. Together with a model of the promoter‐bound state of FrrA our results suggest that inducers act as a wedge, preventing the DNA‐binding domains from moving close enough together to interact with successive positions of the major groove of the palindromic operator.
Article
Full-text available
Soil bacteria from the genus Streptomyces , phylum Actinobacteria, feature a complex metabolism and diverse adaptations to environmental stress. These characteristics are consequences of variable nutrition availability in the soil and allow survival under changing nitrogen conditions. Streptomyces coelicolor is a model organism for Actinobacteria and is able to use nitrogen from a variety of sources including unusual compounds originating from the decomposition of dead plant and animal material, such as polyamines or monoamines (like ethanolamine). Assimilation of nitrogen from these sources in S. coelicolor remains largely unstudied. Using microbiological, biochemical and in silico approaches, it was recently possible to postulate polyamine and monoamine (ethanolamine) utilization pathways in S. coelicolor. Glutamine synthetase-like enzymes (GS-like) play a central role in these pathways. Extensive studies have revealed that these enzymes are able to detoxify polyamines or monoamines and allow the survival of S. coelicolor in soil containing an excess of these compounds. On the other hand, at low concentrations, polyamines and monoamines can be utilized as nitrogen and carbon sources. It has been demonstrated that the first step in poly-/monoamine assimilation is catalyzed by GlnA3 (a γ-glutamylpolyamine synthetase) and GlnA4 (a γ-glutamylethanolamide synthetase), respectively. First insights into the regulation of polyamine and ethanolamine metabolism have revealed that the expression of the glnA3 and the glnA4 gene are controlled on the transcriptional level.
Article
Full-text available
In Streptomyces, antibiotic biosynthesis is triggered in phosphate limitation that is usually correlated with energetic stress. Polyphosphates constitute an important reservoir of phosphate and energy and a better understanding of their role in the regulation of antibiotic biosynthesis is of crucial importance. We previously characterized a gene, SLI_4384/ppk, encoding a polyphosphate kinase, whose disruption greatly enhanced the weak antibiotic production of Streptomyces lividans. In the condition of energetic stress, Ppk utilizes polyP as phosphate and energy donor, to generate ATP from ADP. In this paper, we established that ppk is co-transcribed with its two downstream genes, SLI_4383, encoding a phosin called PptA possessing a CHAD domain constituting a polyphosphate binding module and SLI_4382 encoding a nudix hydrolase. The expression of the ppk/pptA/SLI_4382 operon was shown to be under the positive control of the two-component system PhoR/PhoP and thus mainly expressed in condition of phosphate limitation. However, pptA and SLI_4382 can also be transcribed alone from their own promoter. The deletion of pptA resulted into earlier and stronger actinorhodin production and lower lipid content than the disruption of ppk, whereas the deletion of SLI_4382 had no obvious phenotypical consequences. The disruption of ppk was shown to have a polar effect on the expression of pptA, suggesting that the phenotype of the ppk mutant might be linked, at least in part, to the weak expression of pptA in this strain. Interestingly, the expression of phoR/phoP and that of the genes of the pho regulon involved in phosphate supply or saving were strongly up-regulated in pptA and ppk mutants, revealing that both mutants suffer from phosphate stress. Considering the presence of a polyphosphate binding module in PptA, but absence of similarities between PptA and known exo-polyphosphatases, we proposed that PptA constitutes an accessory factor for exopolyphosphatases or general phosphatases involved in the degradation of polyphosphates into phosphate.
Article
Full-text available
The ATSAS software suite encompasses a number of programs for the processing, visualization, analysis and modelling of small-angle scattering data, with a focus on the data measured from biological macromolecules. Here, new developments in the ATSAS 3.0 package are described. They include IMSIM , for simulating isotropic 2D scattering patterns; IMOP , to perform operations on 2D images and masks; DATRESAMPLE , a method for variance estimation of structural invariants through parametric resampling; DATFT , which computes the pair distance distribution function by a direct Fourier transform of the scattering data; PDDFFIT , to compute the scattering data from a pair distance distribution function, allowing comparison with the experimental data; a new module in DATMW for Bayesian consensus-based concentration-independent molecular weight estimation; DATMIF , an ab initio shape analysis method that optimizes the search model directly against the scattering data; DAMEMB , an application to set up the initial search volume for multiphase modelling of membrane proteins; ELLLIP , to perform quasi-atomistic modelling of liposomes with elliptical shapes; NMATOR , which models conformational changes in nucleic acid structures through normal mode analysis in torsion angle space; DAMMIX , which reconstructs the shape of an unknown intermediate in an evolving system; and LIPMIX and BILMIX , for modelling multilamellar and asymmetric lipid vesicles, respectively. In addition, technical updates were deployed to facilitate maintainability of the package, which include porting the PRIMUS graphical interface to Qt5, updating SASpy – a PyMOL plugin to run a subset of ATSAS tools – to be both Python 2 and 3 compatible, and adding utilities to facilitate mmCIF compatibility in future ATSAS releases. All these features are implemented in ATSAS 3.0 , freely available for academic users at https://www.embl-hamburg.de/biosaxs/software.html.
Article
Full-text available
Enzymes are increasingly applied as biocatalysts for fulfilling industrial needs in a variety of applications and there is a bursting of interest for novel therapeutic proteins. Consequently, developing appropriate expression platforms for efficiently producing such recombinant proteins represents a crucial challenge. It is nowadays widely accepted that an ideal ‘universal microbial host’ for heterologous protein expression does not exist. Indeed, the first-choice microbes, as Escherichia coli or yeasts, possess known intrinsic limitations that inevitably restrict their applications. In this scenario, bacteria belonging to the Streptomyces genus need to be considered with more attention as promising, alternative, and versatile platforms for recombinant protein production. This is due to their peculiar features, first-of-all their natural attitude to secrete proteins in the extracellular milieu. Additionally, streptomycetes are considered robust and scalable industrial strains and a wide range of tools for their genetic manipulation is nowadays available. This mini-review includes an overview of recombinant protein production in streptomycetes, covering nearly 100 cases of heterologous proteins expressed in these Gram-positives from the 1980s to December 2019. We investigated homologous sources, heterologous hosts, and molecular tools (promoters/vectors/signal peptides) used for the expression of these recombinant proteins. We reported on their final cellular localization and yield. Thus, this analysis might represent a useful source of information, showing pros and cons of using streptomycetes as platform for recombinant protein production and paving the way for their more extensive use in future as alternative heterologous hosts.
Article
Full-text available
X‐ray crystallographic analysis of a phosin (PptA) from Steptomyces chartreusis reveals a metal‐associated, lozenge‐shaped fold featuring a 5‐10 Å wide, positively charged tunnel that traverses the protein core. Two distinct metal‐binding sites were identified in which the predominant metal ion was Cu2+. In solution, PptA forms stable homodimers that bind with nanomolar affinity to polyphosphate, a stress‐related biopolymer acting as a phosphate and energy reserve in conditions of nutrient depletion. A single protein dimer interacts with 14‐15 consecutive phosphate moieties within the polymer. Our observations suggest that PptA plays a role in polyphosphate metabolism, mobilisation or sensing, possibly by acting in concert with polyphosphate kinase (Ppk). Like Ppk, phosins may inuence antibiotic synthesis by streptomycetes. This article is protected by copyright. All rights reserved.
Article
Full-text available
In Streptomyces, GlnR is an activator protein that activates nitrogen‐assimilation genes under nitrogen‐limiting conditions. However, less is known regarding the regulation of these genes under nitrogen‐rich conditions. We determined that the developmental regulator MtrA represses nitrogen‐assimilation genes in nitrogen‐rich media and that it competes with GlnR for binding to GlnR boxes. The GlnR boxes upstream of multiple nitrogen genes, such as amtB, were confirmed as MtrA binding sites in vitro by electrophoretic mobility shift assays and in vivo by ChIP‐qPCR analysis. Transcriptional analysis indicated that, on nutrient‐rich medium, MtrA profoundly repressed expression of nitrogen‐associated genes, indicating opposing roles for MtrA and GlnR in the control of nitrogen metabolism. Using in vitro and in vivo analysis, we also showed that glnR is itself a direct target of MtrA and that MtrA represses glnR transcription. We further demonstrated functional conservation of MtrA homologues in the recognition of GlnR boxes upstream of nitrogen genes from different actinobacterial species. As mtrA and glnR are widespread among actinomycetes, this mechanism of potential competitive control over nitrogen metabolism genes may be common in this group, adding a major new layer of complexity to the known regulatory network for nitrogen metabolism in Streptomyces and related species. This article is protected by copyright. All rights reserved.
Article
Full-text available
Most of the phylogenetic diversity of life is found in bacteria and archaea, and is reflected in the diverse metabolism and functions of bacterial and archaeal polyamines. The polyamine spermidine was probably present in the last universal common ancestor, and polyamines are known to be necessary for critical physiological functions in bacteria, such as growth, biofilm formation, and other surface behaviors, and production of natural products such as siderophores. There is also phylogenetic diversity of function, indicated by the role of polyamines in planktonic growth of different species ranging from absolutely essential to entirely dispensable. However, the cellular molecular mechanisms responsible for polyamine function in bacterial growth are almost entirely unknown. In contrast, the molecular mechanisms of essential polyamine functions in archaea are better understood: covalent modification by polyamines of translation factor aIF5A, and the agmatine modification of tRNAIle. As with bacterial hyperthermophiles, archaeal thermophiles require long chain and branched polyamines for growth at high temperatures. For bacterial species in which polyamines are essential for growth, it is still unknown whether the molecular mechanisms underpinning polyamine function involve covalent or noncovalent interactions. Understanding the cellular molecular mechanisms of polyamine function in bacterial growth and physiology remains one of the great challenges for future polyamine research.
Chapter
The exponential growth in the number of newly solved protein structures makes correlating and classifying the data an important task. Distance matrix alignment (Dali) is used routinely by crystallographers worldwide to screen the database of known structures for similarity to newly determined structures. Dali is easily accessible through the web server (http://ekhidna.biocenter.helsinki.fi/dali). Alternatively, the program may be downloaded and pairwise comparisons performed locally on Linux computers.