ArticlePDF Available

Inositol phosphates and core subunits of the Sin3L/Rpd3L histone deacetylase (HDAC) complex up-regulate deacetylase activity

Authors:

Abstract and Figures

The constitutively nuclear histone deacetylases (HDACs) 1, 2, and 3 erase acetyl marks on acetyllysine residues, alter the landscape of histone modifications, and modulate chromatin structure and dynamics and thereby crucially regulate gene transcription in higher eukaryotes. Nuclear HDACs exist as at least six giant, multi-protein complexes whose non-enzymatic subunits confer genome targeting specificity for these enzymes. The deacetylase activity of HDACs previously has been shown to be enhanced by inositol phosphates, which also bridge the catalytic domain in protein-protein interactions with SANT domains in all HDAC complexes except those that contain the Sin3 transcriptional corepressors. Here, using purified recombinant proteins, co-immunoprecipitation and HDAC assays, pulldown and NMR experiments, we show that HDAC1/2 deacetylase activity in one of the most ancient and evolutionarily conserved Sin3L/Rpd3L complexes is inducibly up-regulated by inositol phosphates, but involves interactions with a zinc finger (ZnF) motif in the Sin3-associated protein 30 (SAP30) subunit that is structurally unrelated to SANT domains, indicating convergent evolution at the functional level. This implies that this mode of regulation has independently evolved multiple times and provides an evolutionary advantage. We also found that constitutive association with another core subunit, Rb-binding protein 4 chromatin-binding factor (RBBP4), further enhances deacetylase activity, implying both inducible and constitutive regulatory mechanisms within the same HDAC complex. Our results indicate that inositol phosphates stimulate HDAC activity and that the SAP30 ZnF motif performs roles similar to that of the unrelated SANT domain in promoting the SAP30-HDAC1 interaction and enhancing HDAC activity.
Content may be subject to copyright.
1
Inositol phosphates and core subunits of the Sin3L/Rpd3L histone deacetylase (HDAC) complex up-
regulate deacetylase activity
Ryan Dale Marcum1 and Ishwar Radhakrishnan1,*
From the 1Department of Molecular Biosciences, Northwestern University, 2205 Tech Drive, Evanston,
IL 60208-3500, USA
Running title: Inositol phosphates enhance Sin3L/Rpd3L deacetylase activity
*To whom correspondence should be addressed: Ishwar Radhakrishnan, Department of Molecular
Biosciences, Northwestern University, 2205 Tech Drive, Evanston, IL 60208-3500;
i-radhakrishnan@northwestern.edu; Tel: +1 847-467-1173; Fax: +1 847-467-6489.
Keywords: Transcription regulation, chromatin-modification, histone deacetylase (HDAC), protein-
protein interaction, cofactor-mediated protein-protein interaction, allosteric regulation, inositol phosphate
signaling; convergent evolution; gene expression; epigenetics
ABSTRACT
The constitutively nuclear histone deacetylases
(HDACs) 1, 2, and 3 erase acetyl marks on
acetyllysine residues, alter the landscape of histone
modifications, and modulate chromatin structure
and dynamics and thereby crucially regulate gene
transcription in higher eukaryotes. Nuclear HDACs
exist as at least six giant, multi-protein complexes
whose non-enzymatic subunits confer genome
targeting specificity for these enzymes. The
deacetylase activity of HDACs previously has been
shown to be enhanced by inositol phosphates,
which also bridge the catalytic domain in protein
protein interactions with SANT domains in all
HDAC complexes except those that contain the
Sin3 transcriptional corepressors. Here, using
purified recombinant proteins, co-
immunoprecipitation and HDAC assays, pulldown
and NMR experiments, we show that HDAC1/2
deacetylase activity in one of the most ancient and
evolutionarily conserved Sin3L/Rpd3L complexes
is inducibly up-regulated by inositol phosphates,
but involves interactions with a zinc finger (ZnF)
motif in the Sin3-associated protein 30 (SAP30)
subunit that is structurally unrelated to SANT
domains, indicating convergent evolution at the
functional level. This implies that this mode of
regulation has independently evolved multiple
times and provides an evolutionary advantage. We
also found that constitutive association with another
core subunit, Rb-binding protein 4 chromatin-
binding factor (RBBP4), further enhances
deacetylase activity, implying both inducible and
constitutive regulatory mechanisms within the
same HDAC complex. Our results indicate that
inositol phosphates stimulate HDAC activity and
that the SAP30 ZnF motif performs roles similar to
that of the unrelated SANT domain in promoting
the SAP30HDAC1 interaction and enhancing
HDAC activity.
Lysine acetylation is an abundant post-
translational histone modification found in
euchromatin that is correlated with increased
chromatin dynamics and access to the underlying
DNA template to regulatory factors, effectors, and
molecular machines (1). Histone deacetylation
reverses these effects and the chemical
transformation is mediated, in large part, by
constitutively nuclear, Class I Zn2+-dependent
HDACs 1, 2, and 3 (2,3). In mammals, these
enzymes function in the context of at least six giant,
multiprotein complexes including the
Sin3L/Rpd3L, Sin3S/Rpd3S, NuRD, CoREST,
MiDAC, and SMRT/NCoR complexes that are in
turn recruited via protein-protein interactions by
sequence-specific DNA-binding transcription
factors and/or by post-translational modifications in
the tail regions of histones (4-7). While some
progress has been made regarding how these
complexes are assembled, including the
recruitment and integration of HDACs in these
http://www.jbc.org/cgi/doi/10.1074/jbc.RA119.009780The latest version is at
JBC Papers in Press. Published on July 29, 2019 as Manuscript RA119.009780
by guest on July 29, 2019http://www.jbc.org/Downloaded from
2
complexes, the precise molecular role(s) of
individual subunits has been, with few exceptions,
enigmatic.
The mammalian 1.2-2 megadalton
Sin3L/Rpd3L HDAC complex regulates the
expression of a large variety of genes with roles in
cell cycle control, differentiation, metabolism and
stem cell maintenance (4,7-9). Five subunits of the
complex including Sin3A/B, SAP30/SAP30L,
Sds3/BRMS1/BRMS1L, HDAC1/2 and RBBP4/7
(paralogous subunits are distinguished by a ‘/’) are
found in a broad range of eukaryotes from yeast to
human and are thought to comprise the core
complex (10-14). Among these five subunits,
HDAC1/2 and RBBP4/7 are shared with multiple
chromatin-modifying complexes, Sin3A and Sin3B
assort into the Sin3L/Rpd3L and Sin3S/Rpd3S
complexes, respectively, while the SAP30 and Sds3
subunits (and their paralogs) are unique to the
Sin3L/Rpd3L complex.
We and others have previously shown through
biochemical and/or structural studies that the Sin3A
subunit, besides being directly recruited by
transcription factors via interactions with the PAH1
and PAH2 domains, performs a scaffolding
function by engaging in direct interactions with
HDAC1, Sds3, and SAP30 (4,7). Whereas the
interactions with HDAC1 and Sds3 is mediated by
the HDAC-interaction domain (HID; (15)), the
interaction with SAP30 occurs via the PAH3
domain, located immediately N-terminal to the HID
(16). In these studies, we also showed that Sds3
provides dimerization and putative nucleic acid
binding functions for the complex. However, the
role of SAP30 has been less obvious because,
besides a Sin3-interaction domain (SID), the
polypeptide contains only one other domain that
harbors a unique zinc finger (ZnF) motif that we
and others proposed could bind to negatively-
charged molecules through a highly conserved,
positively-charged surface (17,18). Indeed,
elucidation of the precise molecular roles of the
Sin3, Sds3, and SAP30 polypeptides has been
hampered, due in large part, to the lack of broadly
distributed and well-characterized protein-protein
interaction domains within these proteins.
Recent studies have unexpectedly suggested
that Class I HDACs 1, 2, and 3 are regulated by
inositol phosphates derived from membrane lipids
(19-21). Besides promoting interactions between
HDAC1/3 and SANT-domain bearing proteins via
a conserved surface located near the enzyme active
site, certain inositol phosphates were shown in
these studies to significantly enhance the activity of
the enzyme. However, unlike in the NuRD and
SMRT HDAC complexes, where these inositol
phosphate-dependent interactions have been
described, none of the subunits in the Sin3L/Rpd3L
complex harbor SANT domains. Here, we show
that the SAP30 ZnF motif, which is structurally
unrelated to SANT domains, performs similar roles
in promoting the interaction between SAP30 and
HDAC1 while concomitantly enhancing enzyme
activity. We further show that the histone H3/H4-
interacting subunit RBBP4 also enhances the basal
deacetylase activity of HDAC1 through
constitutive interactions with the catalytic domain.
Results
SAP30 Engages with HDAC1 in the Sin3L/Rpd3L
Complex
Since the SAP30 ZnF motif was previously
shown to bind to molecules enriched in negatively
charged moieties, we surmised that the SAP30
subunit might have a role akin to the SANT
domains of MTA1 and SMRT corepressors in the
NuRD and SMRT complexes (19-21). Specifically,
we asked whether SAP30, and the ZnF motif in
particular, could directly interact with HDAC1. In
co-immunoprecipitation (co-IP) experiments with
HA-tagged HDAC1 and Flag-tagged SAP30, the
wild-type proteins exhibited a robust interaction
(Fig. 1A). However, mutation of C112, a key Zn2+-
coordinating residue, to alanine in SAP30, that we
previously showed precluded proper folding of the
ZnF (17), failed to efficiently immunoprecipitate
HDAC1, implying that a properly folded ZnF motif
was necessary for a robust interaction with
HDAC1. We then asked whether full-length
HDAC1 or a construct lacking a large section of the
unstructured C-terminal tail, but retaining the
catalytic domain, was sufficient for the interaction.
Both constructs were efficiently
immunoprecipitated by SAP30 (Supp. Fig. S1A),
implying that the catalytic domain of HDAC1
served as the primary site of interaction.
To test whether the SAP30 ZnF motif was
sufficient to engage HDAC1 in direct interactions,
we expressed and purified the zinc finger and
conducted pulldown assays with immobilized Flag-
tagged HDAC1. Although no binding was detected
for the His6-tagged SAP30 ZnF when presented
by guest on July 29, 2019http://www.jbc.org/Downloaded from
3
alone (Fig. 1B), inclusion of inositol 1,4,5,6-
tetrakisphosphate (InsP4) resulted in a robust
interaction between the two proteins. However, the
interaction was abrogated when HDAC1 was
presented with SAP30 ZnF pre-treated with EDTA,
even though the binding reaction was supplemented
with InsP4. This again confirmed that the
interaction was specific and that a properly folded
zinc finger was critical for interactions with
HDAC1. Previous studies involving HDAC1/3 and
SANT domains indicated that a minimum of four
phosphate groups at strategic positions in the
inositol ring were necessary for efficient
interactions (21). We tested whether inositol 1,4,5-
triphosphate (InsP3) and inositol 1,2,3,4,5,6-
hexakisphosphate (InsP6) could similarly promote
interactions between the SAP30 ZnF and HDAC1
over a range of inositol phosphate concentrations.
Whereas, InsP6 promoted a robust interaction at
high micromolar concentrations, by contrast, InsP3
did so poorly, even at 100 µM concentration (Fig.
1D). This implies that the same trends noted
previously involving inositol phosphate-mediated
protein-HDAC interactions (21), also apply to the
SAP30-HDAC1 interaction.
We then asked whether SAP30
immunoprecipitates harbored deacetylase activity
in cells. Flag-tagged wild-type SAP30 or the
C112A mutant were co-expressed with HDAC1,
and following immunoprecipitation, eluted from
the anti-Flag resin (Supp. Fig. S1B). Deacetylase
activity assays demonstrated ~4-fold increased
activity in the wild-type immunoprecipitates
compared to the C112A mutant (Fig. 1C).
However, addition of InsP6 did not lead to any
significant increase in deacetylase activity, likely
due to the presence of endogenous InsPs in the
immunoprecipitates. Collectively, these results
demonstrate that SAP30 can directly and
independently recruit deacetylase activity via
interactions with HDAC1 in a manner that is
dependent on an intact ZnF in mammalian cells.
To gain insights into the SAP30 ZnF interaction
with inositol phosphates at the molecular level, we
performed titrations with 15N-labeled protein and
InsP6. The addition of increasing amounts of InsP6
induced relatively small, yet readily discernible
changes in the NMR spectrum with only a subset of
resonances showing significant perturbations (Fig.
1E). The resonances ‘shifted’ as a function of added
ligand, indicative of fast dissociation kinetics for
the complex. Indeed, the modest stability of the
complex was confirmed by the equilibrium
dissociation constant of 250 ± 11 µM obtained
through non-linear least squares fitting of the
binding isotherms (Fig. 1E). Mapping the chemical
shift changes on to the structure of the SAP30 ZnF
motif reveals a contiguous surface with the most
significant perturbations (Supp. Fig. S2). This
surface overlaps with the conserved, basic surface
formed by residues in the α1 and α2 helices that
was previously anticipated to be functionally
relevant (17). However, NMR titration experiments
to probe whether the same surface is also involved
in HDAC1 interactions was precluded by the
limited solubility of the HDAC.
SAP30 ZnF and HDAC1 Interact via Conserved
Surfaces
To determine the role of individual residues in
promoting the formation of the SAP30 ZnF-
HDAC1 complex, we mutated several basic
residues located in the α1 and α2 helices that were
implicated in inositol phosphate binding by our
NMR studies and evaluated them by pulldown
assays. To maximize disruption, the residues were
mutated to glutamic acid either in isolation or in
combination with another residue. Rather
unexpectedly, most of the mutations that were
screened produced only modest or no effects on
binding to HDAC1 (Fig. 2A). The sole exception
was R88E that either in isolation or in combination
with R123E completely abrogated the interaction,
implying that the R88 side chain is most likely
involved in engaging with phosphate moieties in
inositol phosphates.
We then asked whether other residues in the
general vicinity of R88 in the three-dimensional
structure were also important for the interaction. In
this regard, we mutated Q120 and two serines (S84
and S86) in the loop segment preceding helix α1
along with Y110 located in another loop segment
preceding α2 in these interactions. Whereas the
Q120E mutant bound to HDAC1 just like the wild-
type protein, both the S84E,S86E double mutant as
well as the Y110A mutant showed complete loss of
binding to HDAC1 (Fig. 2B). Consistent with the
other results, deletion of the SAP30 ZnF C-terminal
unstructured tail segment (spanning residues 124-
131), which is located distally from R88 and
harbors multiple basic residues that showed
by guest on July 29, 2019http://www.jbc.org/Downloaded from
4
chemical shift perturbations in InsP6 titrations, had
little or no effect on HDAC1-binding activity.
To define the complementary protein surface
involved in the formation of the ternary complex,
we mutated basic residues in HDAC1 that were
previously shown to bind to inositol phosphates
including K31, R270, and R306 to glutamic acid
both individually and in combination (21). The
individual mutations significantly reduced SAP30
ZnF binding activity, while the triple mutant
completely abrogated binding (Fig. 2C). We also
mutated residues near these residues including Y23
and H33 that comprise the binding site in the
MTA1-InsP6-HDAC1 complex. While the Y23A
mutant bound SAP30 ZnF comparably as the wild-
type protein, binding by the H33A mutant was
significantly diminished, implying that residues
beyond those involved in binding inositol
phosphates were important for the stability of the
ternary complex.
Collectively, our mutagenesis studies reveal
that R88, S84, S86, and Y110 side chains in SAP30
ZnF and H33, K31, R270, and R306 in HDAC1 are
critical for the stability of the ternary complex. In
both proteins, the residues define contiguous
surfaces for interactions with inositol phosphates
and HDAC1 (Fig. 2D). To gain more detailed
insights into the mode of interaction involving the
two proteins and InsP6, we performed multi-body
molecular docking of all three components using
HADDOCK with those residues found to be critical
for the interaction designated as ‘active’ residues
for docking (22). The resulting docked structures
were clustered based on the fraction of common
inter-molecular contacts (>0.75). Twenty-eight
structures from the largest cluster (out of 100
structures used for the final refinement in explicit
solvent) with no NOE violations >0.3 Å and
favorable energies were selected for analysis.
While the backbone level precision of the
ensemble for ordered regions is reasonable (0.71 Å
from the average structure), the precision at the side
chain level for interfacial residues is significantly
lower since the restraints were sparse and side
chains at the interface were allowed to move during
the docking (Fig. 3A). Nevertheless, certain
features of the complex could be readily and
consistently detected. InsP6 adopts the chair-like
conformation in previously reported structures and
is sandwiched between the two proteins. One face
of the ‘ring’ of phosphates in InsP6 engages in
electrostatic and/or hydrogen bonding interactions
with K33, R270, and R306 of HDAC1, as noted
previously in the case of the NuRD complex ((21);
Fig. 3B & 3C). Arg88 is the sole side chain in
SAP30 ZnF engaging in salt bridging interactions
with the other face of the phosphate ring. The
hydroxyl groups of S84, S86, and Y110 all engage
different phosphates in hydrogen bonding
interactions. The short side chain of the serine
residues brings the polypeptide backbone in close
proximity to the phosphates, allowing it to engage
in hydrogen bonding interactions. The side chain of
Y110 inserts into a hydrophobic pocket formed by
the side chains of H33, I305, and Y336. Additional
hydrophobic interactions are observed, albeit less
consistently, while other types of favorable
interactions involving interfacial residues are
plausible, even though they are less common in the
ensemble. Notwithstanding these issues, the size of
the protein-protein interface is small, amounting to
a little under 300 Å2, on average, implying that the
primary affinity determinants stem from
interactions with the inositol phosphate moiety.
Indeed, every residue in both proteins involved in
these interactions is invariant over a broad range of
species (Supp. Fig. S3), suggesting strong
evolutionary pressure to preserve these
interactions.
Inositol phosphates, SAP30 and RBBP4 Enhance
HDAC1 Deacetylase Activity
Since certain inositol phosphates enhance
deacetylase activity in other HDAC-containing
multiprotein complexes, we asked whether this was
also the case for the Sin3L/Rpd3L complex. To
answer this question, HDAC1 deacetylase activity
assays using a model acetylated peptide substrate in
end-point format were conducted in the presence of
excess SAP30 ZnF (23). Whereas InsP4, InsP5, and
InsP6 enhanced deacetylase activity two- to three-
fold, albeit at high inositol phosphate
concentrations (>100 µM), InsP3 showed no effect
(Fig. 4A). The deacetylase activity was unaffected
by InsP4 in the absence of SAP30 ZnF. These
results thus imply that a minimum of four
phosphates in inositol phosphates was critical for
the enhancement of HDAC1 deacetylase activity.
To characterize the effects of HDAC1 and
SAP30 ZnF mutations on deacetylase activity, we
measured the kinetic parameters of the enzyme
using the same model substrate above. These assays
by guest on July 29, 2019http://www.jbc.org/Downloaded from
5
were conducted in the presence of 100 µM InsP6.
Whereas the inclusion of SAP30 ZnF only had little
effect on the Km, the kcat was enhanced almost two-
fold (Fig. 4B; Table 1). The HDAC1 R270E mutant
exhibited diminished kcat, although the catalytic
efficiency (kcat/Km ratio) was unchanged from the
wild-type enzyme. The inclusion of SAP30 ZnF led
to only a slight enhancement in both kcat and kcat/Km
for this HDAC1 mutant. Similarly, the SAP30 ZnF
R88E,R123E mutant produced little or no increase
in these parameters relative to that measured for the
enzyme alone. Collectively, these results indicate
that the enhancement of HDAC1 deacetylase
activity emanates from stable formation of the
SAP30 ZnF-InsP6-HDAC1 ternary complex.
Since RBBP4 is known to interact with
HDAC1 constitutively, we asked whether it might
affect HDAC1 deacetylase activity. Compared to
HDAC1 alone, the inclusion of RBBP4 enhanced
kcat almost two-fold, although the Km also increased
significantly to only modestly increase the catalytic
efficiency (Fig. 4C; Table 1). The addition of
SAP30 ZnF further enhanced kcat while diminishing
Km to produce a robust surge in kcat/Km. However,
the effects of RBBP4 and SAP30 ZnF on HDAC1
deacetylase activity appear to be additive rather
than synergistic. In contrast to the enhancement in
deacetylase activity produced by RBBP4 and
SAP30 ZnF individually and in combination,
inclusion of Sin3A HID, which can also associate
constitutively with HDAC1, produced no such
effect (Supp. Fig. S4).
Since enhancements in deacetylase activity
occurred at somewhat high concentrations of
inositol phosphates, presumably reflecting a
modest affinity interaction between SAP30 ZnF
and HDAC1, we asked whether the enhancements
might be more physiologically relevant in the
context of the Sin3L/Rpd3L complex. Since the
interaction between Sin3 and SAP30 is of
nanomolar affinity and the structure of this complex
is known (16), we co-expressed and co-purified a
fusion protein of SAP30 and Sin3A spanning the
ZnF and SID domains in the former and the HID
domain of the latter with HDAC1. Since Sin3A
HID and HDAC1 can associate constitutively, we
conducted activity assays as a function of inositol
phosphate concentration. Whereas InsP3 caused a
modest increase in deacetylase activity at high
concentrations (>100 µM), InsP4, InsP5, and InsP6
produced several-fold enhancements even at low
micromolar concentrations (Fig. 4D; Table 2),
which is well within the intracellular concentrations
measured for these inositol phosphates (24),
implying that these cofactors most likely have a role
in regulating the deacetylase activity of HDAC1.
Discussion
The Sin3L/Rpd3L complex is ancient and one
of the most broadly distributed HDAC complexes
in eukaryotes with half a dozen subunits conserved
from yeast to human (14,25). In yeast, Rpd3, the
ortholog of mammalian HDAC1/2, is found in two
complexes Rpd3L and Rpd3S (26). Although Rpd3
has been reported to be regulated by inositol
phosphates and shares a conserved inositol
phosphate-binding surface with its orthologs from
a variety of species (27), the identity of the subunit
in these HDAC complexes involved in inositol
phosphate-based regulation has remained elusive,
until now. Our studies establish that this role is
performed by the evolutionarily conserved SAP30
subunit in the mammalian Sin3L/Rpd3L complex.
The involvement of the SAP30 zinc finger motif in
these interactions is unexpected because it shares
no overt sequence or structural similarity with
SANT domains that were previously implicated in
this role (19-21). Interestingly, the ZnF motif is
narrowly distributed and is found in only one other
protein in higher eukaryotes that of the SAP30
paralog, SAP30L (28), which is also found in a
subset of Sin3L/Rpd3L complexes, targeting the
complex to the nucleolus (29). The human paralogs
share 79% identity and 89% similarity in this region
with all the residues deemed to be critical for
engaging with inositol phosphates or HDAC1 by
our studies invariant in the two proteins. Thus,
SAP30L is expected to play similar roles as SAP30
in regulating the deacetylase activity of HDACs 1
and 2 in these complexes.
Since inositol phosphate-based activation of
HDAC activity was first described for the human
HDAC3-containing SMRT complex (20), this
paradigm has been extended for the human
HDAC1/2 containing-NuRD (19,21), and now, to
the mammalian Sin3L/Rpd3L complexes (Fig. 5).
Despite differences in the identities of the proteins
involved and the structural motifs used for
engagement with inositol phosphates, each one of
these complexes rely on a minimum of four
phosphate groups in these cofactors for efficient
engagement and activation. Interestingly, whereas
by guest on July 29, 2019http://www.jbc.org/Downloaded from
6
InsP3 has minimal impact on the deacetylase
activity in each of these complexes, InsP4, InsP5,
and InsP6 potentiate the deacetylase activity to
similar extents. The same three basic residues in the
catalytic domain are involved in engaging with the
cofactors in all three complexes, although the
protein-protein interaction surface extends beyond
these residues to varying extents. In all three cases,
engagement with the cofactor and the catalytic
domain results in tangible enhancement in
deacetylase activity. Although a structural basis for
the enhancement is not readily apparent from the
crystal structures, molecular dynamics simulations
have suggested that the effect could be due to
allosteric modulation of protein motions within the
loops comprising the active site (30).
Interestingly, the level of enhancement of
deacetylase activity differs for each complex but is
most critically dependent on the identity of the
HDAC. Whereas the deacetylase activity of
HDAC3 in the SMRT complex is potentiated
substantially by inositol phosphates (~100-fold;
(21)), that of HDAC1 in the NuRD and
Sin3L/Rpd3L complexes is comparably modest (2
to 3-fold; (19)). Although the baseline catalytic rate
and efficiency for HDAC1 is twice as that of
HDAC3, the cofactor-mediated activation for the
latter is much more substantial ((23); note that the
baseline activity of HDAC3 corresponds to when
SMRT is present, since the enzyme is otherwise
inactive). Despite the seemingly modest inositol
phosphate-mediated enhancement in deacetylase
activity for HDAC1, mutations in the inositol
phosphate-binding site causes a substantial
decrease in cell viability (31), highlighting a critical
role played by these small molecule cofactors and
HDAC-associating subunits in multi-protein
complexes.
In all three complexes characterized thus far,
HDACs are recruited and integrated into these
complexes through constitutive associations (by
SMRT, MTA1/2, and Sin3A/B in the SMRT,
NuRD and Sin3L/Rpd3L complexes, respectively),
and therefore, even in the absence of inositol
phosphates, all three complexes retain significant
deacetylase activity. The role of inositol phosphates
thus appears to be in turbocharging the activity of
these enzymes. Since these HDAC complexes have
well-documented roles in regulating the cell cycle
(4-7) and since InsP4, InsP5, and InsP6 levels
fluctuate, peaking in G1 (24), this would suggest
that these cofactors have evolved to function as
facile, molecular signals to tune HDAC activity.
A less appreciated role for inositol phosphates
is that they add another point of contact between the
HDAC and the interacting subunit, contributing to
the tight association and integration of the HDAC
subunits in these complexes. Unlike the MTA1
subunit in the NuRD complex that engages with
HDAC1 both constitutively and inducibly via
structurally distinct domains, association with
HDAC1 in the Sin3L/Rpd3L complex is more
distributed with specific domains of separate
polypeptides Sin3A and SAP30 performing these
roles. Interestingly, RBBP4, which is shared by
both of these complexes and engages in constitutive
interactions with HDAC1 also enhances
deacetylase activity. Structural studies of this
interaction are needed to understand the basis for
this activation.
Its evolutionary conservation notwithstanding,
SAP30 is a subunit unique to the Sin3L/Rpd3L
complex, which raises questions regarding the
identity of the corresponding subunit in the
Sin3S/Rpd3S complex. Since the latter complex
has a different role in transcription elongation in
actively transcribed genes, it is possible that a need
for turbocharging deacetylase activity does not
exist for this complex. Alternatively, since SAP30
ZnF and the SANT domains have independently
evolved to enhance HDAC activity, it is plausible
that additional motifs might yet exist. Indeed,
although SAP30 is evolutionarily conserved, the
SAP30 ortholog in yeast lacks the highly conserved
ZnF motif found in a broad range of organisms
from fly to human (17). Since the inositol
phosphate-binding site is conserved in yeast Rpd3,
we surmise that a yet another structural motif either
within SAP30 or in a different subunit of the Rpd3L
complex likely targets this site, suggesting that this
mechanism of HDAC regulation has independently
evolved multiple times and provides an
evolutionary advantage to the organism.
Experimental Procedures
Protein Expression and Purification
Escherichia coli BL21(DE3) was transformed
with pMCSG7 vector encoding SAP30 ZnF (aa 64-
131) and grown at 37 °C until it reached an OD600nm
of 0.6. Cultures were transferred to 16 °C and
induced with 1 mM isopropyl-β-D-
thiogalactopyranoside for 16 h prior to harvesting.
by guest on July 29, 2019http://www.jbc.org/Downloaded from
7
Cells were suspended in 50 mM Tris-HCl buffer
(pH 8.0) containing 50 mM NaCl, 1 mM Tris-(2-
carboxy-ethyl)-phosphine (TCEP) hydrochloride, 1
mM phenylmethylsulfonyl fluoride (PMSF), 1 µM
leupeptin, 1 mM pepstatin and 0.1% Triton X-100
and lysed by sonication. After centrifugation the
lysate was loaded onto a HiTrap SP cation
exchange column (GE Healthcare). The protein was
eluted using a gradient from 0.05 to 1 M NaCl over
5 column volumes. Fractions were screened by
SDS-PAGE and those containing SAP30 ZnF were
combined and incubated with tobacco etch virus
(TEV) protease overnight to remove the His6 tag at
4 °C; this step was skipped in the case where His6-
tagged protein was required. Further purification
was achieved by reversed phase HPLC using a
linear gradient of 0.1% trifluoracetic acid (TFA)
and 0.1% TFA with 80% acetonitrile. Fractions
were screened by matrix assisted laser
desorption/ionization (MALDI) and lyophilized.
15N labeled samples were purified using the same
method except that cells were grown in 15N
ammonium sulfate containing M9 minimal media.
Point mutations in the SAP30 ZnF expression
construct were introduced using the QuikChange
mutagenesis protocol (Agilent) and the mutants
were produced using the same procedure as the
wild-type protein. All mutations were confirmed by
DNA sequencing.
Full-length human HDAC1 was cloned into a
pcDNA3.1 vector with a C-terminal Flag tag and
TEV protease cleavage site. Flag-tagged HDAC1
was expressed in non-adherent HEK293F cells
following a previously established protocol (32).
Cells were harvested after 48 h of expression and
the cell pellet was resuspended in lysis buffer (50
mM Tris-HCl, pH 7.5, 150 mM NaCl, 50 mM
potassium acetate, 1 mM EDTA, 1 mM PMSF, 1
µM leupeptin, 1 mM pepstatin, 0.2% Triton X-100,
5% glycerol) and sonicated. After centrifugation,
lysates were pre-cleared by incubation with
Sepharose 4B resin (Sigma-Aldrich). Clarified
lysates were incubated with anti-Flag M2 affinity
resin (Sigma-Aldrich) for 1 h at 4 °C. The resin was
then washed with lysis buffer, high salt buffer (50
mM Tris-HCl, pH 7.5, 300 mM NaCl, 50 mM
potassium acetate, 1 mM EDTA, 5% glycerol), and
rinse buffer (50 mM Tris-HCl, pH 7.5, 150 mM
NaCl, 50 mM potassium acetate, 1 mM EDTA),
incubated overnight at 4 °C with TEV protease and
the flow-through containing untagged HDAC1 was
collected. HDAC1 elutions were combined and
dialyzed against 50 mM Tris pH 7.5, 100 mM NaCl,
50 mM KCl, 1 mM TCEP and then 50 mM Tris pH
7.5, 100 mM NaCl, 50 mM KCl, 1 mM TCEP with
10 µM ZnCl2. HDAC1 was co-expressed and co-
purified with RBBP4 using the same protocol as
that for HDAC1 alone.
HA-tagged HDAC1 was co-expressed with a
C-terminally Flag-tagged SAP30 (aa 64-131)-
Sin3A (aa 532-766) fusion construct sub-cloned
into the pcDNA3.1 vector. The complex was
purified by suspending the cell pellet in lysis buffer
and followed by sonication. Lysates were
centrifuged and pre-cleared with Sepharose 4B
resin and incubated with anti-Flag M2 affinity resin
(Sigma-Aldrich) for 1 h at 4 °C. The resin was
washed with lysis, high salt, and rinse buffers after
which it was incubated with 10 µM Sds3 SID (201-
234) for 1 h at 4 °C. The resin was washed with
rinse buffer and the complex was eluted using 150
µg/ml Flag peptide. The sample was immediately
dialyzed in the same manner as HDAC1 alone.
Co-immunoprecipitation Experiments
HEK293T cells were transfected using calcium
phosphate with the indicated expression plasmids
and collected 48 h post-transfection. Cells were
lysed in 20 mM HEPES, pH 7.9, 150 mM KCl, 5%
glycerol, 10 µM zinc acetate, 0.2% NP-40, 1 mM
PMSF, 1 µM leupeptin, 1 mM pepstatin. Cell
extracts were incubated for 1 h at 4 °C with anti-
Flag M2 affinity resin (Sigma-Aldrich). The resin
was washed five times with high-salt buffer (20
mM HEPES, pH 7.9, 300 mM KCl, 5% glycerol, 10
µM zinc acetate, 0.2% NP-40) and samples were
boiled with SDS-PAGE loading buffer. Proteins
were then resolved using SDS-PAGE, transferred
to nitrocellulose, and then probed with primary
antibodies, either anti-Flag (Sigma-Aldrich,
#F3165, 1:500 dilution) or anti-HA (Sigma-
Aldrich, #H3663, 1:500 dilution) after which anti-
mouse HRP-conjugated secondary antibody
(Thermo Fisher Scientific, #OB617005, 1:1000
dilution) was used. The blot was imaged using West
Pico chemiluminescent substrate (Thermo
Scientific, #34080) and a Syngene Pxi
chemiluminescent imager. Mutations in the Flag-
tagged, full-length human SAP30 and HA-tagged
human HDAC1 constructs were introduced using
the QuikChange protocol (Agilent); all mutations
were confirmed by DNA sequencing. Wild-type
by guest on July 29, 2019http://www.jbc.org/Downloaded from
8
and mutant SAP30-HDAC1 complexes were
immunoprecipitated and eluted using excess Flag
peptide (ApexBio) for deacetylase activity assays.
Pulldown Experiments
HDAC1 was expressed and purified as
described above except that it was left bound to the
anti-Flag M2 affinity resin. SAP30 ZnF wild-type
and mutants were expressed as previously
described and pellets were resuspended in pulldown
buffer (20 mM HEPES, pH 7.9, 100 mM NaCl, 50
mM potassium acetate, 5% glycerol, 0.1% NP-40,
1 mM PMSF, 1 µM leupeptin, 1 mM pepstatin) and
lysed by sonication. Cell lysates were incubated
with HDAC1-bound anti-Flag resin in the presence
of 100 µM InsP4 or InsP6 for 30 min. The resin was
washed five times with pulldown buffer and boiled
in SDS-PAGE loading buffer. Samples were
resolved by SDS-PAGE, transferred to
nitrocellulose, stained with either anti-Flag or anti-
His (Thermo Fisher, MA121315, 1:1000 dilution)
antibody. Blots were visualized in the same manner
as described above for the co-IP experiments.
NMR Experiments
NMR data were acquired on a 600 MHz Agilent
DD2 NMR spectrometer at 25 °C. Dry, lyophilized
15N-SAP30 ZnF was dissolved in 50 mM Tris-d11
acetate-d4 buffer (pH 6.0) and 0.2% NaN3 with
equimolar ZnCl2 added. The inositol phosphates
InsP4 (Cayman Chemicals) and InsP6 (Sigma-
Aldrich) used for the titrations were dissolved in the
same buffer and used without additional
purification. Data processing and analysis was
performed using Felix (Felix NMR) and
NMRFAM-Sparky (33).
HDAC1 Deacetylase Activity Assays
A model peptide substrate containing an
aminocoumarin (AMC) moiety was used for the
deacetylase assays. The Ac-Gly-Ala-Lys(Ac)-
AMC peptide was purchased from Bachem
Americas and used in a slightly modified protocol
described previously (23). Reactions were
performed in a 384-well plate at room temperature
with 1 nM HDAC1 in 50 mM HEPES, pH 7.4, 100
mM KCl, 0.001% Tween-20, 5% DMSO, and 200
nM trypsin. Fluorescence data were acquired at
room temperature for 1 h on a Biotek Synergy 4
microplate reader with excitation and emission
wavelengths set to 360 and 440 nm, respectively.
Data were fitted with R using non-linear least
squares fitting. Endpoint assays were conducted in
a similar manner as above but using a fixed
concentration of 60 µM of the AMC peptide.
Steady-state fluorescence in these assays was
measured after 1 h of equilibration.
Molecular Modeling
The previously determined NMR structure of
SAP30 ZnF (PDB code: 2KDP; (17)) was docked
with the crystal structure of HDAC1 bound to InsP6
(PDB code: 5ICN; (21)) using HADDOCK version
2.2 (22). Active residues for docking were assigned
based on the results of pulldown experiments with
mutant proteins. These included S84, S86, R88, and
Y110 in SAP30 ZnF and K31, H33, R270, and
R306 in HDAC1, besides the InsP6 moiety. To
sample a broader range of conformations, the first
ten conformers of SAP30 ZnF in the NMR
ensemble were used in the calculations. Force field
parameters for InsP6 were generated for CNS using
the PRODRG server (34); these were supplemented
with additional chirality restraints. One thousand
structures were calculated during the rigid body
docking phase; the 200 best structures were used for
semiflexible docking, of which the 100 best
structures were refined in explicit solvent. Default
parameters were used except the distance restraints
were weighted more significantly (0.1, 0.5, 1.0 for
the three phases of the docking calculation).
Distance restraints between specific InsP6
phosphate groups and K31, R270, and R306 to
target distances observed in the crystal structure
were enforced throughout the calculation to ensure
the InsP6 moiety was anchored to HDAC1 (21).
Analysis of the resulting structures were conducted
using PLIP (35), LIGPLOT (36), and MONSTER
(37).
Acknowledgements: We would like to thank Gregory David (New York University) for critical reading of
this manuscript. We are grateful to the Robert Lurie Comprehensive Cancer Center at Northwestern for
supporting structural biology research.
by guest on July 29, 2019http://www.jbc.org/Downloaded from
9
Conflicts of interest: The authors declare that they have no conflicts of interest with the contents of this
article.
by guest on July 29, 2019http://www.jbc.org/Downloaded from
10
References
1. Bannister, A. J., and Kouzarides, T. (2011) Regulation of chromatin by histone modifications.
Cell Res. 21, 381-395
2. Seto, E., and Yoshida, M. (2014) Erasers of histone acetylation: the histone deacetylase enzymes.
Cold Spring Harb Perspect Biol 6, a018713
3. Yang, X. J., and Seto, E. (2008) The Rpd3/Hda1 family of lysine deacetylases: from bacteria and
yeast to mice and men. Nat. Rev. Mol. Cell Biol. 9, 206-218
4. Adams, G. E., Chandru, A., and Cowley, S. M. (2018) Co-repressor, co-activator and general
transcription factor: the many faces of the Sin3 histone deacetylase (HDAC) complex. Biochem.
J. 475, 3921-3932
5. Hayakawa, T., and Nakayama, J. (2011) Physiological roles of class I HDAC complex and
histone demethylase. Journal of biomedicine & biotechnology 2011, 129383
6. Watson, P. J., Fairall, L., and Schwabe, J. W. (2012) Nuclear hormone receptor co-repressors:
structure and function. Mol. Cell. Endocrinol. 348, 440-449
7. Grzenda, A., Lomberk, G., Zhang, J. S., and Urrutia, R. (2009) Sin3: master scaffold and
transcriptional corepressor. Biochim. Biophys. Acta 1789, 443-450
8. Silverstein, R. A., and Ekwall, K. (2005) Sin3: a flexible regulator of global gene expression and
genome stability. Curr. Genet. 47, 1-17
9. Chaubal, A., and Pile, L. A. (2018) Same agent, different messages: insight into transcriptional
regulation by SIN3 isoforms. Epigenetics Chromatin 11, 17
10. David, G., Turner, G. M., Yao, Y., Protopopov, A., and DePinho, R. A. (2003) mSin3-associated
protein, mSds3, is essential for pericentric heterochromatin formation and chromosome
segregation in mammalian cells. Genes Dev. 17, 2396-2405
11. Fleischer, T. C., Yun, U. J., and Ayer, D. E. (2003) Identification and characterization of three
new components of the mSin3A corepressor complex. Mol. Cell. Biol. 23, 3456-3467
12. McDonel, P., Costello, I., and Hendrich, B. (2009) Keeping things quiet: roles of NuRD and Sin3
co-repressor complexes during mammalian development. Int. J. Biochem. Cell Biol. 41, 108-116
13. Bowman, C. J., Ayer, D. E., and Dynlacht, B. D. (2014) Foxk proteins repress the initiation of
starvation-induced atrophy and autophagy programs. Nat. Cell Biol. 16, 1202-1214
14. Banks, C. A. S., Thornton, J. L., Eubanks, C. G., Adams, M. K., Miah, S., Boanca, G., Liu, X.,
Katt, M. L., Parmely, T. J., Florens, L., and Washburn, M. P. (2018) A Structured Workflow for
Mapping Human Sin3 Histone Deacetylase Complex Interactions Using Halo-MudPIT Affinity-
Purification Mass Spectrometry. Mol. Cell. Proteomics 17, 1432-1447
15. Clark, M. D., Marcum, R., Graveline, R., Chan, C. W., Xie, T., Chen, Z., Ding, Y., Zhang, Y.,
Mondragon, A., David, G., and Radhakrishnan, I. (2015) Structural insights into the assembly of
the histone deacetylase-associated Sin3L/Rpd3L corepressor complex. Proc. Natl. Acad. Sci. U.
S. A. 112, E3669-3678
16. Xie, T., He, Y., Korkeamaki, H., Zhang, Y., Imhoff, R., Lohi, O., and Radhakrishnan, I. (2011)
Structure of the 30-kDa Sin3-associated protein (SAP30) in complex with the mammalian Sin3A
corepressor and its role in nucleic acid binding. J. Biol. Chem. 286, 27814-27824
17. He, Y., Imhoff, R., Sahu, A., and Radhakrishnan, I. (2009) Solution structure of a novel zinc
finger motif in the SAP30 polypeptide of the Sin3 corepressor complex and its potential role in
nucleic acid recognition. Nucleic Acids Res. 37, 2142-2152
18. Viiri, K. M., Janis, J., Siggers, T., Heinonen, T. Y., Valjakka, J., Bulyk, M. L., Maki, M., and
Lohi, O. (2009) DNA-binding and -bending activities of SAP30L and SAP30 are mediated by a
zinc-dependent module and monophosphoinositides. Mol. Cell. Biol. 29, 342-356
19. Millard, C. J., Watson, P. J., Celardo, I., Gordiyenko, Y., Cowley, S. M., Robinson, C. V., Fairall,
L., and Schwabe, J. W. (2013) Class I HDACs Share a Common Mechanism of Regulation by
Inositol Phosphates. Mol. Cell 51, 57-67
by guest on July 29, 2019http://www.jbc.org/Downloaded from
11
20. Watson, P. J., Fairall, L., Santos, G. M., and Schwabe, J. W. (2012) Structure of HDAC3 bound
to co-repressor and inositol tetraphosphate. Nature 481, 335-340
21. Watson, P. J., Millard, C. J., Riley, A. M., Robertson, N. S., Wright, L. C., Godage, H. Y.,
Cowley, S. M., Jamieson, A. G., Potter, B. V., and Schwabe, J. W. (2016) Insights into the
activation mechanism of class I HDAC complexes by inositol phosphates. Nature
communications 7, 11262
22. Dominguez, C., Boelens, R., and Bonvin, A. M. (2003) HADDOCK: a protein-protein docking
approach based on biochemical or biophysical information. J. Am. Chem. Soc. 125, 1731-1737
23. Schultz, B. E., Misialek, S., Wu, J., Tang, J., Conn, M. T., Tahilramani, R., and Wong, L. (2004)
Kinetics and comparative reactivity of human class I and class IIb histone deacetylases.
Biochemistry 43, 11083-11091
24. Barker, C. J., Wright, J., Hughes, P. J., Kirk, C. J., and Michell, R. H. (2004) Complex changes in
cellular inositol phosphate complement accompany transit through the cell cycle. Biochem. J.
380, 465-473
25. Carrozza, M. J., Florens, L., Swanson, S. K., Shia, W. J., Anderson, S., Yates, J., Washburn, M.
P., and Workman, J. L. (2005) Stable incorporation of sequence specific repressors Ash1 and
Ume6 into the Rpd3L complex. Biochim. Biophys. Acta 1731, 77-87; discussion 75-76
26. Carrozza, M. J., Li, B., Florens, L., Suganuma, T., Swanson, S. K., Lee, K. K., Shia, W. J.,
Anderson, S., Yates, J., Washburn, M. P., and Workman, J. L. (2005) Histone H3 methylation by
Set2 directs deacetylation of coding regions by Rpd3S to suppress spurious intragenic
transcription. Cell 123, 581-592
27. Worley, J., Luo, X., and Capaldi, A. P. (2013) Inositol pyrophosphates regulate cell growth and
the environmental stress response by activating the HDAC Rpd3L. Cell Rep 3, 1476-1482
28. Laitaoja, M., Tossavainen, H., Pihlajamaa, T., Valjakka, J., Viiri, K., Lohi, O., Permi, P., and
Janis, J. (2016) Redox-dependent disulfide bond formation in SAP30L corepressor protein:
Implications for structure and function. Protein Sci. 25, 572-586
29. Viiri, K. M., Korkeamaki, H., Kukkonen, M. K., Nieminen, L. K., Lindfors, K., Peterson, P.,
Maki, M., Kainulainen, H., and Lohi, O. (2006) SAP30L interacts with members of the Sin3A
corepressor complex and targets Sin3A to the nucleolus. Nucleic Acids Res. 34, 3288-3298
30. Arrar, M., Turnham, R., Pierce, L., de Oliveira, C. A., and McCammon, J. A. (2013) Structural
insight into the separate roles of inositol tetraphosphate and deacetylase-activating domain in
activation of histone deacetylase 3. Protein Sci. 22, 83-92
31. Jamaladdin, S., Kelly, R. D., O'Regan, L., Dovey, O. M., Hodson, G. E., Millard, C. J., Portolano,
N., Fry, A. M., Schwabe, J. W., and Cowley, S. M. (2014) Histone deacetylase (HDAC) 1 and 2
are essential for accurate cell division and the pluripotency of embryonic stem cells. Proc. Natl.
Acad. Sci. U. S. A. 111, 9840-9845
32. Portolano, N., Watson, P. J., Fairall, L., Millard, C. J., Milano, C. P., Song, Y., Cowley, S. M.,
and Schwabe, J. W. (2014) Recombinant protein expression for structural biology in HEK 293F
suspension cells: a novel and accessible approach. J. Vis. Exp., e51897
33. Lee, W., Tonelli, M., and Markley, J. L. (2015) NMRFAM-SPARKY: enhanced software for
biomolecular NMR spectroscopy. Bioinformatics 31, 1325-1327
34. Schuttelkopf, A. W., and van Aalten, D. M. (2004) PRODRG: a tool for high-throughput
crystallography of protein-ligand complexes. Acta Crystallogr. D Biol. Crystallogr. 60, 1355-
1363
35. Salentin, S., Schreiber, S., Haupt, V. J., Adasme, M. F., and Schroeder, M. (2015) PLIP: fully
automated protein-ligand interaction profiler. Nucleic Acids Res. 43, W443-447
36. Wallace, A. C., Laskowski, R. A., and Thornton, J. M. (1995) LIGPLOT: a program to generate
schematic diagrams of protein-ligand interactions. Protein Eng. 8, 127-134
37. Salerno, W. J., Seaver, S. M., Armstrong, B. R., and Radhakrishnan, I. (2004) MONSTER:
inferring non-covalent interactions in macromolecular structures from atomic coordinate data.
Nucleic Acids Res. 32, W566-568
by guest on July 29, 2019http://www.jbc.org/Downloaded from
12
FOOTNOTES
This work was supported by grants from the American Heart Association (14GRNT20170003 and
17GRNT33680167) to I.R. R.D.M. was supported by pre-doctoral fellowships from the National Institute
of General Medical Sciences (T32 GM008382) and the American Heart Association (16PRE27260041).
The abbreviations used are: HDAC, histone deacetylase; HID, HDAC interaction domain; SID, Sin3
interaction domain; InsP, inositol phosphate; co-IP, co-immunoprecipitation; TCEP, Tris-(2-carboxy-
ethyl)-phosphine; PMSF, phenylmethylsulfonyl fluoride; TEV, tobacco etch virus; MALDI, matrix assisted
laser desorption/ionization; TFA, trifluoroacetic acid; AMC, aminocoumarin.
by guest on July 29, 2019http://www.jbc.org/Downloaded from
13
Table 1. Kinetic parameters for wild-type and mutant HDAC1 enzyme activity*
Protein(s)
Km, µM
kcat, s-1
kcat/Km, M-1 s-1
HDAC1
HDAC1+SAP30 ZnF$
HDAC1 R270E
HDAC1 R270E+SAP30 ZnF$
HDAC1+SAP30 ZnF R88E,R123E$
HDAC1+RBBP4
HDAC1+RBBP4+SAP30 ZnF
$
13.98 ± 1.08
12.37 ± 0.68
10.24 ± 0.70
9.65 ± 1.25
17.36 ± 3.93
18.92 ± 1.63
11.78 ± 1.62
2.88 ± 0.07
4.75 ± 0.08
2.12 ± 0.05
2.46 ± 0.11
3.19 ± 0.24
4.89 ± 0.14
6.35 ± 0.25
2.06 × 105
3.84 × 105
2.07 × 105
2.55 × 105
1.84 × 105
2.58 × 105
5.39 × 10
5
*uncertainties correspond to 95% confidence intervals
$experiments with SAP30 ZnF proteins conducted in the presence of 30 µM ZnF and 100 µM D-myo-
inositol 1,2,3,4,5,6-hexakisphosphate (InsP6)
Table 2. EC50 parameters for HDAC1 activation by inositol phosphates*
Inositol phosphate
EC50, µM
D-myo-inositol 1,4,5-triphosphate (InsP
3
)
D-myo-inositol 1,4,5,6-tetrakisphosphate (InsP4)
D-myo-inositol 1,3,4,5,6-pentakisphosphate (InsP5)
D-myo-inositol 1,2,3,4,5,6-hexakisphosphate (InsP6)
n.d.
5.18 ± 1.03
5.19 ± 0.88
6.07 ± 0.57
*uncertainties correspond to standard deviations from three independent measurements
by guest on July 29, 2019http://www.jbc.org/Downloaded from
14
Figure 1. SAP30 physically associates via the ZnF domain with the catalytic domain of HDAC1. (A)
Domain map of SAP30. Co-immunoprecipitation (Co-IP) of HA-tagged HDAC1 by Flag-tagged, wild-type
and the C112A mutant of SAP30. (B) Pulldown assays conducted with immobilized Flag-tagged HDAC1
and His6-tagged SAP30 ZnF in the presence or absence of D-myo-inositol 1,4,5,6-tetrakisphosphate (InsP4)
and EDTA. (C) Endpoint activity assays with wild-type and C112A SAP30 immunoprecipitates, with and
without the addition of 100 µM InsP6.; error bars represent standard deviations (n=3). (D) Pulldown assays
conducted as in panel C, except in the presence or absence of varying amounts of InsP3 or InsP6. (E) 1H-
15N correlated spectrum of a 250 µM sample of 15N-SAP30 ZnF in the absence (magenta) and presence
(cyan) of eight equivalents of InsP6 (left panel). Expanded plot showing changes in peak position for V122
in the same titration (top right) and binding isotherms deduced from changes in chemical shifts for selected
residues (bottom right). In panels A, B, and D, molecular weights (in KDa) are shown on the right-hand
side of each blot.
by guest on July 29, 2019http://www.jbc.org/Downloaded from
15
Figure 2. Mapping residues at the interfaces of the SAP30 ZnF-InsP6-HDAC1 complex via mutagenesis
and interaction assays. Pulldown assays conducted with immobilized Flag-tagged HDAC1 and His6-tagged
SAP30 ZnF single-site mutants of candidate basic residues involved in engaging with inositol phosphates
(A) or other solvent-exposed conserved residues (B). The protein labeled 124-131 corresponds to SAP30
ZnF that has a C-terminal deletion; the construct spans residues 64-123. (C) Pulldown assays with
immobilized Flag-tagged wild-type or mutant HDAC1 and His6-tagged SAP30 ZnF. All pulldown
experiments were conducted in the presence of InsP6. (D) Mutations mapped onto the surface of SAP30
ZnF (right) and HDAC1 (left), purple indicates mutations that lead to loss of binding in pulldown
experiments and green indicates no change was observed. In panels A, B, and C, molecular weights (in
KDa) are shown on the right-hand side of each blot.
by guest on July 29, 2019http://www.jbc.org/Downloaded from
16
Figure 3. A structural model for the SAP30 ZnF-InsP6-HDAC1 complex obtained via molecular docking.
SAP30 ZnF backbone and side chain carbon atoms are colored in cyan while HDAC1 backbone and side
chain carbon and InsP6 carbon atoms are colored in tan; non-carbon atoms are shown in Corey-Pauling-
Koltun (CPK) colors. (A) The ensemble of 28 docked structures deemed to be consistent with the functional
studies described in Figure 2. (B) A close-up view of the interface showing the side chains of interacting
residues in the representative structure of the ensemble. (C) Stabilizing, inter-molecular hydrogen bonding
interactions (green lines) involving the side chains of essential residues in the two proteins and InsP6 in the
representative model in 3D (left panel) and in 2D (right panel). Dashed lines in the latter representation
indicate hydrogen bonds while semicircles with radiating spokes identify residues involved in inter-
molecular hydrophobic contacts. Carbon atoms in the 2D representation are colored in black while the bond
colors are in cyan for SAP30 ZnF, tan for HDAC1, and orange for InsP6.
by guest on July 29, 2019http://www.jbc.org/Downloaded from
17
Figure 4. Enhancement in the catalytic activity of HDAC1 by SAP30 and RBBP4. (A) Changes in the level
of HDAC1 deacetylase activity measured in an end-point assay format as a function of inositol phosphate
concentration in the presence of 30 µM SAP30 ZnF. Changes in activity are reported relative to the level
of activity measured in the absence of inositol phosphates. The data shown in gray and denoted ‘InsP4 (
SAP30 ZnF)’ measured deacetylase activity in the presence of InsP4 but in the absence of SAP30 ZnF. (B)
Deacetylase assays conducted using wild-type or mutant SAP30 ZnF and/or wild-type or mutant HDAC1
in the presence of 100 µM InsP6. (C) Changes in the kinetic parameters in deacetylase assays of HDAC1
conducted in the absence or presence of RBBP4, SAP30 ZnF and 100 µM InsP6. (D) Changes in the level
of deacetylase activity measured in an end-point assay format as a function of inositol phosphate
concentration in the presence of stoichiometric amounts of HDAC1 and SAP30-Sin3A fusion protein.
Changes in activity are reported relative to the level of activity measured in the absence of inositol
phosphates.
by guest on July 29, 2019http://www.jbc.org/Downloaded from
18
Figure 5. Structural comparison of various HDAC complexes. (A) Crystal structure of the MTA1-InsP6-
HDAC1 complex (PDB accession code: 5ICN), (B) structural model of the SAP30 ZnF-InsP6-HDAC1
complex from molecular docking, and (C) crystal structure of the SMRT DAD-InsP4-HDAC3 complex
(PDB accession code: 4A69). Note the parallels in InsP and HDAC engagement by the MTA1 and SMRT
subunits featuring structurally similar SANT domains, whereas the mode of InsP and HDAC engagement
by the structurally unrelated SAP30 ZnF domain is unique, implying convergent evolution at the functional
level.
by guest on July 29, 2019http://www.jbc.org/Downloaded from
Ryan Dale Marcum and Ishwar Radhakrishnan
(HDAC) complex up-regulate deacetylase activity
Inositol phosphates and core subunits of the Sin3L/Rpd3L histone deacetylase
published online July 29, 2019J. Biol. Chem.
10.1074/jbc.RA119.009780Access the most updated version of this article at doi:
Alerts:
When a correction for this article is posted When this article is cited
to choose from all of JBC's e-mail alertsClick here
by guest on July 29, 2019http://www.jbc.org/Downloaded from
... Mutation of the conserved IP-binding residues abolished HDAC3 enzyme activity in vitro 11 and deceased HDAC1/2 activity in living cells 12 . All Class I HDACs tested have activated enzyme kinetics in the presence of higher-order phosphorylated species of inositol phosphates 6,7,13 . However, the extent to which Class 1 HDACs require inositol phosphates in vivo remains unclear. ...
Preprint
Histone deacetylases (HDACs) repress transcription by catalyzing the removal of acetyl groups from histones. Class 1 HDACs are activated by inositol phosphate signaling molecules in vitro , but it is unclear if this regulation occurs in human cells. Inositol Polyphosphate Multikinase (IPMK) is required for production of inositol hexakisphosphate (IP6), pentakisphosphate (IP5) and certain tetrakisphosphate (IP4) species, all known activators of Class 1 HDACs in vitro . Here, we generated IPMK knockout (IKO) human U251 glioblastoma cells, which decreased cellular inositol phosphate levels and increased histone H4-acetylation by mass spectrometry. ChIP-seq showed IKO increased H4-acetylation at IKO-upregulated genes, but H4-acetylation was unchanged at IKO-downregulated genes, suggesting gene-specific responses to IPMK knockout. HDAC deacetylase enzyme activity was decreased in HDAC3 immunoprecipitates from IKO vs . wild-type cells, while deacetylase activity of other Class 1 HDACs had no detectable changes in activity. Wild-type IPMK expression in IKO cells fully rescued HDAC3 deacetylase activity, while kinase-dead IPMK expression had no effect. Further, the deficiency in HDAC3 activity in immunoprecipitates from IKO cells could be fully rescued by addition of synthesized IP4 (Ins(1,4,5,6)P4) to the enzyme assay, while control inositol had no effect. These data suggest that cellular IPMK-dependent inositol phosphates are required for full HDAC3 enzyme activity and proper histone H4-acetylation. Implications for targeting IPMK in HDAC3-dependent diseases are discussed.
... HMTs are effector enzymes functioning as methyl group transfer catalysts from methyl donor S-adenosyl methionine (SAM) to arginine and lysine residues in protein targeting classes. Evidence shows that about 70 enzymes are catalysts for histone amino acid residue methylation [173]. The first HMT, Su (var) 3-9 (SUV39H1), was identified in humans and mice. ...
Article
Full-text available
Purpose of Review Breast cancer is a metastatic carcinoma that is recognized as a leading cause of death in women worldwide and is still an unanswered medical challenge. Epigenetics involve reversible changes in the gene expression, without variations in the gene sequence, and mainly include histone post-translational modifications and DNA methylation. The central therapeutics used for cancer, like surgery, radiation, chemotherapy, hormonal therapy, and immunotherapy, have several side effects that raise questions about their use. This suggests that there is an imperative need to seek a safer alternative that has fewer or minimal side effects, and naturally occurring phytochemicals emerge as a potential candidate for cancer treatment by exhibiting anti-neoplastic properties by acting on post-translational histone modifications. In this review article, we discussed the beneficial effects of phytochemicals (isolated or mixed) on breast cancer–associated post-translational histone modifications involved in changing chromatin structure and regulating the transcriptional activity of specific oncogenes which are crucial for causing breast cancer. Recent Findings Preclinical studies have shown that the use of natural phytochemicals causes cell cycle arrest, inhibits signal transduction, and exhibits anti-cancer properties. Furthermore, the phytochemicals studied in this review article were found to modulate the expression and epigenetic activity of histone acetyltransferases, Histone deacetylases (HDACs), and methyltransferases. Summary The results obtained from published in vitro and in vivo models of breast cancer help scientists to develop novel therapeutic drugs to target signaling pathways and reduce the mortality index worldwide. However, there is still a need to conduct more clinical trials using naturally occurring phytochemicals to explore their potential in regulating epigenetic events in cancer. Graphical Abstract Beneficial effects of dietary phytochemicals in regulating histone modifications and treating BC: the figure highlights the role of phytochemicals in regulating histone post-translational modifications by regulating activities of histone-modifying enzymes and thus preventing alteration in gene expression and breast cancer progression.
... Deacetylase assays were performed using a model acetylated peptide substrate as previously described 64 . The concentration of the Rpd3L complex used in these assays were estimated based on SDS-PAGE band intensities of various subunits following Coomassie Brilliant Blue staining compared to known amounts of bovine serum albumin; band intensities were quantified using AzureSpot software (Azure Biosystems). ...
Article
Full-text available
The Rpd3L histone deacetylase (HDAC) complex is an ancient 12-subunit complex conserved in a broad range of eukaryotes that performs localized deacetylation at or near sites of recruitment by DNA-bound factors. Here we describe the cryo-EM structure of this prototypical HDAC complex that is characterized by as many as seven subunits performing scaffolding roles for the tight integration of the only catalytic subunit, Rpd3. The principal scaffolding protein, Sin3, along with Rpd3 and the histone chaperone, Ume1, are present in two copies, with each copy organized into separate lobes of an asymmetric dimeric molecular assembly. The active site of one Rpd3 is completely occluded by a leucine side chain of Rxt2, while the tips of the two lobes and the more peripherally associated subunits exhibit varying levels of flexibility and positional disorder. The structure reveals unexpected structural homology/analogy between unrelated subunits in the fungal and mammalian complexes and provides a foundation for deeper interrogations of structure, biology, and mechanism of these complexes, as well as for the discovery of HDAC complex-specific inhibitors.
... Additionally, we identified ZmSANT gene. SANT domain protein was reported to be associated with chromatin remodeling, histone acetylation and deacetylation, but the biological function is unknown (Boyer et al., 2004;Marcum and Radhakrishnan, 2019). Fox system is reliable and efficient for screening up-regulated genes under salt stress or other abiotic stresses. ...
Article
Full-text available
Zoysia matrella is a salt-tolerant turfgrass grown in areas with high soil salinity irrigated with effluent water. Previous studies focused on explaining the regulatory mechanism of Z. matrella salt-tolerance at phenotypic and physiological levels. However, the molecular mechanism associated with salt tolerance of Z. matrella remained unclear. In this study, a high-efficient method named FOX (full-length cDNA overexpression) hunting system was used to search for salt-tolerant genes in Z. matrella. Eleven candidate genes, including several known or novel salt-tolerant genes involved in different metabolism pathways, were identified. These genes exhibited inducible expression under salt stress condition. Furthermore, a novel salt-inducible candidate gene ZmGnTL was transformed into Arabidopsis for functional analysis. ZmGnTL improved salt-tolerance through regulating ion homeostasis, reactive oxygen species scavenging, and osmotic adjustment. In summary, we demonstrated that FOX is a reliable system for discovering novel genes relevant to salt tolerance and several candidate genes were identified from Z. matrella that can assist molecular breeding for plant salt-tolerance improvement.
... In the previous structures, InsP6 was involved in the structure stabilization, ternary interactions, and folding [25,26]. InsP6 was also reported to play a role in the RNA editing [25], mRNA transcription [27], RNA export [28], and DNA repair [29,30] and the regulation of the histone deacetylases (HDAC) activity [31]. InsP6 also acts as a glue by bringing Cullin-RING ligase (CRL) and COP9 signalosome (CSN) together, and plays a role in UV radiation resistance [32]. ...
Article
Full-text available
Proteasomes comprise a family of proteasomal complexes essential for maintaining protein homeostasis. Accordingly, proteasomes represent promising therapeutic targets in multiple human diseases. Several proteasome inhibitors are approved for treating hematological cancers. However, their side effects impede their efficacy and broader therapeutic applications. Therefore, understanding the biology of the different proteasome complexes present in the cell is crucial for developing tailormade inhibitors against specific proteasome complexes. Here, we will discuss the structure, biology, and function of the alternative Proteasome Activator 200 (PA200), also known as PSME4, and summarize the current evidence for its dysregulation in different human diseases. We hereby aim to stimulate research on this enigmatic proteasome regulator that has the potential to serve as a therapeutic target in cancer.
... Initially, IP4 was found to be an intrinsic scaffolding part located at a binding pocket formed at the interface between the HDACs and their cognate corepressors NuRD, CoREST, SMRT, Sin3L/Rpd3L and others (82)(83)(84)(85)(86)(87). It was then demonstrated that IP4 allosterically activates the enzymatic activity of the HDACs which are not functional in its absence and outside of the corepressor complexes (88)(89)(90). Interestingly, IP5 and IP6 can substitute for IP4 at least in vitro (90). ...
Article
Full-text available
Chromatin, the complex of DNA and histone proteins, serves as a main integrator of cellular signals. Increasing evidence links cellular functional to chromatin state. Indeed, different metabolites are emerging as modulators of chromatin function and structure. Alterations in chromatin state are decisive for regulating all aspects of genome function and ultimately have the potential to produce phenotypic changes. Several metabolites such as acetyl-CoA, S-adenosyl methionine (SAM) or adenosine triphosphate (ATP) have now been well characterized as main substrates or cofactors of chromatin modifying enzymes. However, there are other metabolites that can directly interact with chromatin influencing its state or that modulate the properties of chromatin regulatory factors. Also, there is a growing list of atypical enzymatic and non-enzymatic chromatin modifications that originate from different cellular pathways that have not been in the limelight of chromatin research. Here, we summarize different properties and functions of uncommon regulatory molecules originating from intermediate metabolism of lipids, carbohydrates and amino acids. Based on the various modes of action on chromatin and the plethora of putative, so far not described chromatin regulating metabolites, we propose that there are more links between cellular functional state and chromatin regulation to be discovered. We hypothesize that these connections could provide interesting starting points for interfering with cellular epigenetic states at a molecular level.
Article
Full-text available
The interplay between genetic alterations and metabolic dysregulation is increasingly recognized as a pivotal axis in cancer pathogenesis. Both elements are mutually reinforcing, thereby expediting the ontogeny and progression of malignant neoplasms. Intriguingly, recent findings have highlighted the translocation of metabolites and metabolic enzymes from the cytoplasm into the nuclear compartment, where they appear to be intimately associated with tumor cell proliferation. Despite these advancements, significant gaps persist in our understanding of their specific roles within the nuclear milieu, their modulatory effects on gene transcription and cellular proliferation, and the intricacies of their coordination with the genomic landscape. In this comprehensive review, we endeavor to elucidate the regulatory landscape of metabolic signaling within the nuclear domain, namely nuclear metabolic signaling involving metabolites and metabolic enzymes. We explore the roles and molecular mechanisms through which metabolic flux and enzymatic activity impact critical nuclear processes, including epigenetic modulation, DNA damage repair, and gene expression regulation. In conclusion, we underscore the paramount significance of nuclear metabolic signaling in cancer biology and enumerate potential therapeutic targets, associated pharmacological interventions, and implications for clinical applications. Importantly, these emergent findings not only augment our conceptual understanding of tumoral metabolism but also herald the potential for innovative therapeutic paradigms targeting the metabolism–genome transcriptional axis.
Article
Full-text available
Lysine acetylation in histone tails is a key post-translational modification that controls transcription activation. Histone deacetylase complexes remove histone acetylation, thereby repressing transcription and regulating the transcriptional output of each gene. Although these complexes are drug targets and crucial regulators of organismal physiology, their structure and mechanisms of action are largely unclear. Here, we present the structure of a complete human SIN3B histone deacetylase holo-complex with and without a substrate mimic. Remarkably, SIN3B encircles the deacetylase and contacts its allosteric basic patch thereby stimulating catalysis. A SIN3B loop inserts into the catalytic tunnel, rearranges to accommodate the acetyl-lysine moiety, and stabilises the substrate for specific deacetylation, which is guided by a substrate receptor subunit. Our findings provide a model of specificity for a main transcriptional regulator conserved from yeast to human and a resource of protein-protein interactions for future drug designs.
Article
Full-text available
The SET3 complex (SET3C) is a seven-subunit histone deacetylase complex that is capable of transcriptional regulation. Methylated histone 3 marks recruit SET3C to the nucleosome, and the SET3C catalytic subunits deacetylate the histone 3 and 4 tails. There is very limited structural knowledge of the SET3C subunits, with most subunits having unknown structures or functions. Here, a catalytically active SET3 complex was endogenously purified from Saccharo­myces cerevisiae and utilized for negative-stain electron microscopy (EM) to determine an apo model for the holo complex. The negative-stain EM 3D model revealed a three-lobe architecture, with each lobe extending from a central point.
Article
Full-text available
At face value, the Sin3 histone deacetylase (HDAC) complex appears to be a prototypical co-repressor complex, that is, a multi-protein complex recruited to chromatin by DNA bound repressor proteins to facilitate local histone deacetylation and transcriptional repression. While this is almost certainly part of its role, Sin3 stubbornly refuses to be pigeon-holed in quite this way. Genome-wide mapping studies have found that Sin3 localises predominantly to the promoters of actively transcribed genes. While Sin3 knockout studies in various species result in a combination of both up- and down-regulated genes. Furthermore, genes such as the stem cell factor, Nanog, are dependent on the direct association of Sin3 for active transcription to occur. Sin3 appears to have properties of a co-repressor, co-activator and general transcription factor, and has thus been termed a co-regulator complex. Through a series of unique domains, Sin3 is able to assemble HDAC1/2, chromatin adaptors and transcription factors in a series of functionally and compositionally distinct complexes to modify chromatin at both gene-specific and global levels. Unsurprisingly, therefore, Sin3/HDAC1 have been implicated in the regulation of numerous cellular processes, including mammalian development, maintenance of pluripotency, cell cycle regulation and diseases such as cancer.
Article
Full-text available
SIN3 is a global transcriptional coregulator that governs expression of a large repertoire of gene targets. It is an important player in gene regulation, which can repress or activate diverse gene targets in a context-dependent manner. SIN3 is required for several vital biological processes such as cell proliferation, energy metabolism, organ development, and cellular senescence. The functional flexibility of SIN3 arises from its ability to interact with a large variety of partners through protein interaction domains that are conserved across species, ranging from yeast to mammals. Several isoforms of SIN3 are present in these different species that can perform common and specialized functions through interactions with distinct enzymes and DNA-binding partners. Although SIN3 has been well studied due to its wide-ranging functions and highly conserved interaction domains, precise roles of individual SIN3 isoforms have received less attention. In this review, we discuss the differences in structure and function of distinct SIN3 isoforms and provide possible avenues to understand the complete picture of regulation by SIN3.
Article
Full-text available
Although a variety of affinity purification mass spectrometry (AP-MS) strategies have been used to investigate complex interactions, many of these are susceptible to artifacts due to substantial overexpression of the exogenously expressed bait protein. Here we present a logical and systematic workflow that uses the multifunctional Halo tag to assess the correct localization and behavior of tagged subunits of the Sin3 histone deacetylase complex prior to further AP-MS analysis. Using this workflow, we modified our tagging/expression strategy with 21.7% of the tagged bait proteins that we constructed, allowing us to quickly develop validated reagents. Specifically, we apply the workflow to map interactions between stably expressed versions of the Sin3 subunits SUDS3, SAP30 or SAP30L and other cellular proteins. Here we show that the SAP30 and SAP30L paralogues strongly associate with the core Sin3 complex, but SAP30L has unique associations with the proteasome and the myelin sheath. Next, we demonstrate an advancement of the complex NSAF (cNSAF) approach, in which normalization to the scaffold protein SIN3A accounts for variations in the proportion of each bait capturing Sin3 complexes and allows a comparison between different baits capturing the same protein complex. This analysis reveals that although the Sin3 subunit SUDS3 appears to be used in both SIN3A and SIN3B based complexes, the SAP30 subunit is not used in SIN3B based complexes. Intriguingly, we do not detect the Sin3 subunits SAP18 and SAP25 among the 128 high-confidence interactions identified, suggesting that these subunits may not be common to all versions of the Sin3 complex in human cells. This workflow provides the framework for building validated reagents to assemble quantitative interaction networks for chromatin remodeling complexes and provides novel insights into focused protein interaction networks.
Article
Full-text available
Histone deacetylases (HDACs) 1, 2 and 3 form the catalytic subunit of several large transcriptional repression complexes. Unexpectedly, the enzymatic activity of HDACs in these complexes has been shown to be regulated by inositol phosphates, which bind in a pocket sandwiched between the HDAC and co-repressor proteins. However, the actual mechanism of activation remains poorly understood. Here we have elucidated the stereochemical requirements for binding and activation by inositol phosphates, demonstrating that activation requires three adjacent phosphate groups and that other positions on the inositol ring can tolerate bulky substituents. We also demonstrate that there is allosteric communication between the inositol-binding site and the active site. The crystal structure of the HDAC1:MTA1 complex bound to a novel peptide-based inhibitor and to inositol hexaphosphate suggests a molecular basis of substrate recognition, and an entropically driven allosteric mechanism of activation.
Article
Full-text available
Significance Gene transcription in eukaryotes is regulated by enzymes that posttranslationally add or remove acetyl groups from histones and render the underlying DNA more or less accessible to the transcription machinery. How histone deacetylases (HDACs), the enzymes responsible for deacetylation that are commonly found in multiprotein complexes, are assembled and targeted to their sites of action to affect transcription repression is largely unknown. We show biochemically and structurally how two key subunits of a conserved HDAC complex recruit multiple copies of HDACs into the complex in a manner that allows the enzymes to explore a large conformational space when the complex is targeted to specific genomic loci. This complex seems to be tailored for efficient deacetylation of nucleosomes that are situated far apart.
Article
Full-text available
The characterization of interactions in protein-ligand complexes is essential for research in structural bioinformatics, drug discovery and biology. However, comprehensive tools are not freely available to the research community. Here, we present the protein-ligand interaction profiler (PLIP), a novel web service for fully automated detection and visualization of relevant non-covalent protein-ligand contacts in 3D structures, freely available at projects.biotec.tu-dresden.de/plip-web. The input is either a Protein Data Bank structure, a protein or ligand name, or a custom protein-ligand complex (e.g. from docking). In contrast to other tools, the rule-based PLIP algorithm does not require any structure preparation. It returns a list of detected interactions on single atom level, covering seven interaction types (hydrogen bonds, hydrophobic contacts, pi-stacking, pi-cation interactions, salt bridges, water bridges and halogen bonds). PLIP stands out by offering publication-ready images, PyMOL session files to generate custom images and parsable result files to facilitate successive data processing. The full python source code is available for download on the website. PLIP's command-line mode allows for high-throughput interaction profiling. © The Author(s) 2015. Published by Oxford University Press on behalf of Nucleic Acids Research.
Article
Full-text available
SPARKY (Goddard & Kneller, SPARKY 3) remains the most popular software program for NMR data analysis, despite the fact that development of the package by its originators ceased in 2001. We have taken over the development of this package and describe NMRFAM-SPARKY, which implements new functions reflecting advances in the biomolecular NMR field. NMRFAM-SPARKY has been repackaged with current versions of Python and Tcl/Tk, which support new tools for NMR peak simulation and graphical assignment determination. These tools, along with chemical shift predictions from the PACSY database (Lee et al. 2012), greatly accelerate protein side chain assignments. NMRFAM-SPARKY supports automated data format interconversion for interfacing with a variety of web-servers including, PECAN (Eghbalnia et al. 2005), PINE (Bahrami et al. 2009), TALOS-N (Shen and Bax 2013), CS-Rosetta (Shen et al. 2009), SHIFTX2 (Han et al. 2011) and PONDEROSA-C/S (Lee et al. 2014). Availability: The software package, along with binary and source codes, if desired, can be downloaded freely from http://pine.nmrfam.wisc.edu/download_packages.html. Instruction manuals and video tutorials can be found at http://www.nmrfam.wisc.edu/nmrfam-sparky-distribution.htm. whlee@nmrfam.wisc.edu; markley@nmrfam.wisc.edu. © The Author(s) 2014. Published by Oxford University Press.
Article
Sin3A-associated protein 30-like (SAP30L) is one of the key proteins in a multi-subunit protein complex involved in transcriptional regulation via histone deacetylation. SAP30L, together with a highly homologous SAP30 as well as other SAP proteins (i.e., SAP25, SAP45, SAP130 and SAP180), is an essential component of the Sin3A corepressor complex, although its actual role has remained elusive. SAP30L is thought to function as an important stabilizing and bridging molecule in the complex and to mediate its interactions with other corepressors. SAP30L has been previously shown to contain an N-terminal Cys3 His type zinc finger (ZnF) motif, which is responsible for the key protein-protein, protein-DNA and protein-lipid interactions. By using high-resolution mass spectrometry, we studied a redox-dependent disulfide bond formation in SAP30L ZnF as a regulatory mechanism for its structure and function. We showed that upon oxidative stress SAP30L undergoes the formation of two specific disulfide bonds, a vicinal Cys29-Cys30 and Cys38-Cys74, with a concomitant release of the coordinated zinc ion. The oxidized protein was shown to remain folded in solution and to bind signaling phospholipids. We also determined a solution NMR structure for SAP30L ZnF that showed an overall fold similar to that of SAP30, determined earlier. The NMR titration experiments with lipids and DNA showed that the binding is mediated by the C-terminal tail as well as both α-helices of SAP30L ZnF. The implications of these results for the structure and function of SAP30L are discussed. This article is protected by copyright. All rights reserved.
Article
The LIGPLOT program automatically generates schematic 2-D representations of protein-ligand complexes from standard Protein Data Bank file input. The output is a colour, or black-and-white, PostScript file giving a simple and informative representation of the intermolecular interactions and their strengths, including hydrogen bonds, hydrophobic interactions and atom accessibilities. The program is completely general for any ligand and can also be used to show other types of interaction in proteins and nucleic acids. It was designed to facilitate the rapid inspection of many enzyme complexes, but has found many other applications.