Sie sind auf Seite 1von 13

Biochimica et Biophysica Acta 1789 (2009) 45–57

Contents lists available at ScienceDirect

Biochimica et Biophysica Acta


j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / b b a g r m

Review

Chemical mechanisms of histone lysine and arginine modifications


Brian C. Smith a, John M. Denu b,⁎
a
Department of Chemistry, University of Wisconsin-Madison, Madison, Wisconsin 53706, USA
b
Department of Biomolecular Chemistry, University of Wisconsin-Madison, 1300 University Avenue, Madison, Wisconsin 53706-1532, USA

a r t i c l e i n f o a b s t r a c t

Article history: Histone lysine and arginine residues are subject to a wide array of post-translational modifications
Received 1 May 2008 including methylation, citrullination, acetylation, ubiquitination, and sumoylation. The combinatorial
Accepted 9 June 2008 action of these modifications regulates critical DNA processes including replication, repair, and
Available online 14 June 2008
transcription. In addition, enzymes that modify histone lysine and arginine residues have been correlated
with a variety of human diseases including arthritis, cancer, heart disease, diabetes, and neurodegenerative
Keywords:
Methyltransferase
disorders. Thus, it is important to fully understand the detailed kinetic and chemical mechanisms of these
Acetyltransferase enzymes. Here, we review recent progress towards determining the mechanisms of histone lysine and
Demethylase arginine modifying enzymes. In particular, the mechanisms of S-adenosyl-methionine (AdoMet) dependent
Deacetylase methyltransferases, FAD-dependent demethylases, iron dependent demethylases, acetyl-CoA dependent
Deiminase acetyltransferases, zinc dependent deacetylases, NAD+ dependent deacetylases, and protein arginine
Sirtuin deiminases are covered. Particular attention is paid to the conserved active-site residues necessary for
catalysis and the individual chemical steps along the catalytic pathway. When appropriate, areas requiring
further work are discussed.
© 2008 Elsevier B.V. All rights reserved.

1. Introduction ification of lysine and arginine residues within histone proteins.


However, it is important to point out that similar reactions on non-
Within eukaryotic organisms, the basic unit of chromosomes is the histone proteins will proceed through the same mechanism. We will
nucleosome which is composed of double-stranded DNA wrapped pay particular attention to the conserved active-site residues
around a protein octamer containing two copies each of the histone necessary for catalysis and the individual chemical steps along the
proteins H2A, H2B, H3, and H4 [1,2]. Histone proteins are subject to a catalytic pathway. Where possible we will try to incorporate both
wide array of post-translational modifications including methylation, the former and new nomenclature of these enzymes designated by
citrullination (deimination), acetylation, phosphorylation, ubiquitina- former name/new name (for explanation of the new nomenclature
tion, and sumoylation occurring within the histone core region as well see [10]).
as on the N-terminal tails that protrude from the core region [3]. The
combinatorial influence of these modifications in both time and space 2. Lysine modifying enzymes
affects important DNA regulatory processes including replication,
repair, and transcription [3]. Lysine residues within histones are subject to a variety of
Within histone proteins, lysine and arginine residues are modifications including methylation, acetylation, ubiquitination,
abundant and highly post-translationally modified. Enzymes that and sumoylation on their ɛ-amino groups [3]. Of the enzymes that
modify these lysine and arginine residues have been correlated with catalyze histone post-translational modification, those that modify
a variety of human disease states such as rheumatoid arthritis [4], lysine are the best understood. In general, histone lysine acetylation
cancer [5], heart disease [6], diabetes [7,8], as well as neurodegen- is correlated with gene activation whereas deacetylation is corre-
erative disorders such as Parkinson's disease and Alzheimer's disease lated with gene repression/silencing [11]. However, histone lysine
[8,9]. In light of the importance of these enzymes in a large variety of methylation is correlated with either activation or repression
human disease states, it is critical to elucidate their catalytic depending on the site and degree of methylation. Given their
mechanisms. important involvement in gene regulation, it is not surprising that
In this review, we will focus on the kinetic and chemical the enzymes that regulate these post-translational modifications are
mechanisms of the enzymes that perform post-translational mod- implicated in a variety of human disease states. For example, histone
lysine methyltransferases are implicated in a wide variety of cancers
⁎ Corresponding author. Tel.: +1 608 265 1859; fax: +1 608 262 5253. [12], histone acetyltransferases are associated with leukemia [13],
E-mail address: jmdenu@wisc.edu (J.M. Denu). aberrant recruitment of class I/II histone deacetylases (HDACs) are

1874-9399/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.bbagrm.2008.06.005
46 B.C. Smith, J.M. Denu / Biochimica et Biophysica Acta 1789 (2009) 45–57

linked to leukemias and lymphomas [14], and class III HDACs are that neither Tyr-335 nor any other active-site residue are positioned to
involved in age-associated diseases such as type II diabetes, obesity, act as a general base [33,34] and theoretical studies calculate a pKa for
Parkinson's disease, and Alzheimer's disease [7–9]. Therefore, Tyr-335 of N13.0 [35]. Alternatively, calculations reveal that placing
elucidating the chemical mechanisms catalyzed by these enzymes the positive charge of the protonated ɛ-amine in closely proximity to
will be important therapeutically. In fact, a small-molecule inhibitor the cationic methylsulfonium lowers the pKa of the ɛ-amine from
of class I/II HDACs was recently FDA-approved for the treatment of ≥10.9 to ≤8.2 in Set7/9 [35,36]. As processive mechanisms have been
cutaneous T-cell lymphoma [14]. proposed [17,26,29,37] for the formation of dimethyl- and trimethyl-
lysine in some SET domain methyltransferases, multiple rounds of
2.1. Histone lysine methyltransferases lysine deprotonation must occur without release of methylated lysine.
It has been hypothesized that an ordered water channel created upon
Lysine methyltransferases catalyze mono-, di-, and trimethylation formation of the ternary complex between Set7/9, the unmodified or
of the lysine ɛ-amino group in an S-adenosyl-L-methionine (Ado- methylated lysine substrate, and AdoMet allows release of protons to
Met) dependent manner. Histone lysine methyltransferases consist bulk solvent [35,36].
of two main classes, the SET domain containing family and the DOT1 For multiple methylations to occur processively, the lysine ɛ-
family. carbon–nitrogen bond must rotate so that the ɛ-amine can be further
deprotonated and the resulting lone pair aligned with the methyl–
2.1.1. SET domain histone lysine methyltransferases sulfur bond of AdoMet. In the monomethyl specific methyltransferase,
SET domains are found in a few bacterial proteins and a large Set7/9, Tyr-245 and Tyr-305 hydrogen bond with the lysine substrate
number of eukaryotic proteins with over 60 human members [15] to prevent rotation of the ɛ-carbon–nitrogen bond and therefore
although not all have known histone lysine methyltransferase activity. alignment of the lone pair with the methyl–sulfur bond of AdoMet
SET domain methyltransferases exist which terminally form mono-, [18,23,26,35,38]. Mutation of Tyr-245 or Tyr-305 of Set7/9 to
di-, or trimethyl-lysine. The size and bonding patterns of the active- phenylalanine converts the enzyme to a tri- or dimethyltransferase,
site residues determine whether the enzyme carries out mono-, di-, or respectively [26,38]. Interestingly, in the SET domain methyltrans-
trimethylation of its target lysine residue. ferases DIM-5 and G9a/KMT1C [10], which are capable of trimethyla-
SET domain methyltransferases catalyze a sequential bi–bi tion and dimethylation, respectively, the equivalent residue to Tyr-305
kinetic mechanism in which both substrate association and product is a phenylalanine. Therefore, this tyrosine to phenylalanine substitu-
release occur in a random fashion [16,17]. Within the active-site, a tion is thought to act as a switch to catalyze methylation beyond
conserved aspartate or glutamate hydrogen bonds to the two-ribose monomethylation.
hydroxyls of AdoMet. In addition, a conserved arginine or lysine Several other means for the selective formation of distinct
(Lys-294 in Set7/9/hKMT7 [10]) forms a salt bridge with the methylation states have been hypothesized. In Rubisco large subunit
carboxylate of AdoMet. These interactions position AdoMet in a U- methyltransferase (LSMT), which forms trimethyl-lysine, it is thought
shaped conformation that places the methylsulfonium group at the that a more spacious lysine-binding pocket accommodates the
base of a hydrophobic channel in which the lysine substrate binds. increasing bulk of monomethyl- and dimethyl-lysine within the
In Set7/9, the side chain hydroxyls of Tyr-245 and Tyr-305 hydrogen active-site [39]. Alternatively, it has been proposed from several
bond to the ɛ-amine of lysine, directing the lone pair towards the theoretical studies that formation of an ordered water channel (see
methyl group of AdoMet [18]. The AdoMet methyl group is above) is responsible for catalysis of a specific methylation state. In
positioned and activated by C–H…O hydrogen bonds between the particular, it was proposed that methyl-lysine substrate binding to
methyl group and the side chain hydroxyl of Tyr-335 and the main Set7/9 blocks this water channel so that the bound monomethyl-
chain carboxyls of Gly-264 and His-293. The energy of C–H…O lysine cannot be deprotonated, consistent with rate-limiting depro-
hydrogen bonds is estimated to be 0.5–3 kcal/mol, which is 2–10- tonation and a calculated pKa of 13.8 for protonated monomethyl-
fold weaker than more classical N/O–H…O hydrogen bonds (see [19– lysine [36]. In contrast, a separate theoretical study on Set7/9
21] and references therein). However, in folded proteins the indicated that methyl transfer, not lysine deprotonation, is rate-
contribution of C–H…O hydrogen bonds may be more significant limiting and that formation of monomethyl-lysine creates an ordered
due to a greater desolvation penalty for placing N/O–H…O compared water channel [35,36]. This channel results in a hydrogen bond
to C–H…O hydrogen bonds within a hydrophobic protein environ- between the lone pair of monomethyl-lysine nitrogen and the nearest
ment [22]. Within SET domain methyltransferases these C–H…O water molecule in the channel, which does not allow the lone pair to
hydrogen bonds position the two substrates such that the sulfur of react with AdoMet in a subsequent methylation step [40]. Both of
AdoMet, the carbon of the transferred methyl, and the ɛ-nitrogen of these mechanisms account for the selective formation of mono-
lysine are collinear. This geometry allows SN2 nucleophilic attack of methyl-lysine products by Set7/9, but future work is necessary to
the ɛ-amine to transfer the methyl group from AdoMet and form distinguish between these mechanisms.
the products monomethyl-lysine and S-adenosyl-L-homocysteine
(AdoHcy) (Fig. 1) [23]. After one round of methyl transfer, it is 2.1.2. Dot1/KMT4 histone lysine methyltransferases
hypothesized that C–H…O hydrogen-bonding between the methy- Dot1/KMT4 [10] histone lysine methyltransferases do not contain a
lammonium group and active-site residues orient monomethyl- or SET domain and methylate Lys-79 within the core domain of histone
dimethyl-lysine either for subsequent methylation events or to H3 [41–44]. In addition, Dot1/KMT4 methyltransferases only methy-
disfavor subsequent methyl transfer [24]. late nucleosomal substrates, not free histones. Structures of Dot1/
For methyl transfer to occur, the ɛ-amine of the lysine substrate KMT4 methyltransferases with AdoMet or AdoHcy bound reveal an
must be deprotonated. Several mechanisms of deprotonation have extended AdoMet conformation that is in stark contrast to the folded
been proposed including deprotonation by an active-site residue (e.g. U-shaped conformation found in SET domain methyltransferases
Tyr-245 [25,26] or Tyr-335 [18,25,27,28] in Set7/9), an ordered water [45,46]. Sequence analysis suggests that Dot1/KMT4 methyltrans-
molecule [29], bulk solvent [30], or deprotonation by decreasing the ɛ- ferases possess AdoMet binding motifs similar to histone arginine
amine pKa through close proximity to the cationic methylsulfonium methyltransferase family [47].
group of AdoMet [31]. Alternatively, the observation that SET domain Despite lacking a SET domain, Dot1/KMT4 methyltransferases
methyltransferases possess an unusually high pH optimum of ∼10 catalyze an overall similar mechanism of methyl transfer (Fig. 1).
suggests that the lysine substrate might be deprotonated upon Critical contacts between AdoMet and Thr-139, Gln-168, Glu-186,
binding [32]. However, structural and mutagenesis studies indicate Asp-161, and Asp-222 of human DOT1L/hKMT4 position the
B.C. Smith, J.M. Denu / Biochimica et Biophysica Acta 1789 (2009) 45–57 47

Fig. 1. Proposed general chemical mechanism of AdoMet-dependent histone lysine methyltransferases. SET domain, Dot1/KMT4, and protein arginine methyltransferases use a
similar mechanism of methyl transfer.

methylsulfonium group at the end of a channel in which the lysine LSD1/KDM1 catalyzes the concurrent reduction of FAD to FADH2
substrate binds [45,46]. Similar to SET domain methyltransferases, and oxidation of the methylated lysine substrate generating an imine
this channel aligns the methyl group of AdoMet with the ɛ-nitrogen intermediate (Fig. 2). The formation of this imine requires a free lone
of lysine for a SN2 methyl transfer reaction [45]. Also similar to SET pair on lysine, and accordingly, LSD1/KDM1 is only capable of
methyltransferases, the AdoMet and lysine substrate binding sites demethylating mono- and dimethyl-lysine residues. While no
are on separate faces of the enzyme indicating that the lysine structures of methylated lysine peptides bound to LSD1/KDM1 exist,
substrate can undergo multiple rounds of methylation without modeling of monomethyl-lysine from a methionine residue bound
being released from the enzyme, consistent with a processive within the LSD1/KDM1 active-site predicts a distance of ∼ 3 Å from the
mechanism. However, the fact that a mixture of unmodified, mono-, methyl group (the site of oxidative demethylation) to the reactive N-5
di-, and trimethylated Lys-79 exists in yeast suggests that methyl atom of FAD [49]. Once the imine is formed, water is thought to non-
transfer proceeds distributively [44]. enzymatically attack this intermediate to form a hemiaminal that
As no residues capable of deprotonating lysine are present in the subsequently collapses to formaldehyde and amine. Although other
active-site of Dot1/KMT4 methyltransferases, several hypotheses have electron acceptors have been postulated to exist [50], the FADH2 in
been proposed for lysine deprotonation. One hypothesis is that the LSD1/KDM1 is generally thought to be reoxidized to FAD by one
carboxylate of AdoMet acts as a general base to deprotonate lysine equivalent of molecular oxygen to stoichiometrically produce hydro-
[45]. Alternatively, it has been proposed that close proximity of the ɛ- gen peroxide. FADH2 oxidation is likely rapid as the optical spectrum
amine to the positively charged methylsulfonium and/or the overall of LSD1/KDM1 reveals absorbance maxima at ∼ 370 and 458 nm in the
hydrophobic environment of the active-site lower the pKa of the lysine resting state [51], indicative of fully oxidized FAD [52].
residue [32]. Regardless of the mechanism of lysine deprotonation, the The precise mechanism of imine formation in LSD1/KDM1 is
significantly different pH profiles between Dot1/KMT4 and SET controversial and may occur via hydride transfer or single electron
domain methyltransferases suggests that Dot1 methyltransferases mechanisms [53–55]. A recent study of an FAD-dependent oxidase,
use a distinct mechanism of lysine deprotonation for AdoMet attack tryptophan 2-monooxygenase, is informative. In this study, the
[28,39,46]. Future studies may utilize this mechanistic difference in authors measured a large deuterium isotope effect (6.0 ± 0.5) for
the design specific inhibitors of human DOT1L/hKMT4. In addition, carbon–hydrogen bond cleavage in the conversion of amine to imine
future work is needed to understand the mechanism of specific [54]. This large deuterium isotope effect is consistent with irreversible
methylation of a histone core residue, Lys-79, in favor of the more carbon–hydrogen bond cleavage, which favors a hydride transfer over
exposed lysine residues located on the N-terminal histone tails within a radical mechanism. Whether this is true for LSD1/KDM1 will require
nucleosomal substrates. further investigation.

2.2. Lysine specific histone demethylases 2.2.2. JHDM family


The discovery of LSD1/KDM1 opened the door towards finding
2.2.1. LSD1/KDM1 family other demethylases, particularly those capable of demethylating
Until recently, it was unclear if methylation of lysine residues was a trimethyl-lysine residues, an activity that LSD1/KDM1 does not
permanent modification. In 1973, Paik and Kim were the first to observe possess. In 2006, Tsukada et al. identified the first [56] of a large
demethylation activity of calf thymus histones, but were unable to family of histone demethylases that each contains a jumonji C domain
isolate the active enzyme [48]. It took another three decades for the first critical for demethylase activity. These enzymes have thus been
histone demethylase to be isolated and characterized. In 2004, a lysine- designated jumonji histone demethylases (JHDMs). These enzymes
specific demethylase, LSD1/KDM1 [10], was shown to demethylate are conserved from yeast to humans with approximately 30 members
mono- and dimethylated Lys-4 of histone H3 with concomitant identified in humans [57].
formation of formaldehyde. LSD1/KDM1 belongs to the amine oxidase JHDMs belongs to the Fe(II)/2-oxoglutarate-dependent dioxygen-
superfamily and oxidatively demethylates lysine residues in a FAD- ase (or hydroxylase) superfamily [58], which catalyze a diverse set of
dependent manner (for a more detailed discussion see [53]). reactions via Fe(IV)-oxo species [59]. In the proposed JHDM
48 B.C. Smith, J.M. Denu / Biochimica et Biophysica Acta 1789 (2009) 45–57

Fig. 2. Proposed chemical mechanism of LSD1/KDM1. This figure was adapted from Culhane and Cole [53].

mechanism, a quaternary complex containing 2-oxoglutarate, Fe(II), intermediate oxidizes the methyl carbon of the methylated lysine
and methylated lysine substrate reacts with molecular oxygen (Fig. 3). producing a hemiaminal intermediate and regenerating Fe(II). As with
In the active-site, the conserved residues His-188, Glu-190, and His- LSD1/KDM1, this hemiaminal is thought to spontaneously decompose
276 (JMJD2A/KDM4A numbering [10]) chelate the Fe(II) [60–62]. In to demethylated lysine and formaldehyde. In contrast with LSD1/
the first chemical step, an electron is transferred from Fe(II) to KDM1, the mechanism employed by JHDM enzymes does not require a
molecular oxygen to generate a superoxide radical and Fe(III) [57,63]. lone electron pair on the ɛ-nitrogen of the methylated lysine substrate
It is hypothesized that the hydrophobic active-site assists to generate allowing demethylation of trimethyl-lysine substrates.
high iron oxidation states [57]. Nucleophilic attack of the activated Recent structures of JMJD2A/KDM4A [60–62], which catalyzes
oxygen at the ketone carbon of 2-oxoglutarate results in an Fe(IV) demethylation of di- and trimethylated Lys-9 and trimethylated Lys-
peroxyhemiketal bicyclic intermediate [58]. Subsequent decarboxyla- 36 of histone H3, provide important insights into the selectivity of
tion produces succinate, CO2, and an Fe(IV)-oxo intermediate. This these enzymes for certain methylation states. In JMJD2A/KDM4A, the

Fig. 3. Proposed chemical mechanism of JHDM enzymes.


B.C. Smith, J.M. Denu / Biochimica et Biophysica Acta 1789 (2009) 45–57 49

methylammonium binding pocket is composed of the carbonyl 2.3. Histone acetyltransferases


oxygen of Gly-170 and the side chains of Tyr-177, Glu-190, Ser-288,
and Asn-290. These residues are hypothesized to participate in C–H…O Histone acetyltransferases (HATs or KATs [10]) catalyze the transfer
hydrogen-bonding [19] with the polarized methyl group(s) of of the acetyl moiety from acetyl-CoA to the ɛ-amino of histone lysine
methylated lysine residues [60]. When the substrate is trimethylated, residues resulting in acetylated lysine and CoA. HATs are grouped into
these C–H…O hydrogen bonds position one of the methyl groups three main families: GNAT (for its founding member yeast Gcn5/
towards the iron center where this methyl group can undergo ScKAT2 [10]), MYST (for the founding members MOZ, Ybf2/Sas3, Sas2,
oxidation. In contrast, consistent with the slow demethylation rates and Tip60), and p300/CBP [70].
of JMJD2A/KDM4A with monomethyl- and dimethyl- compared to
trimethyl-lysine substrates, these C–H…O hydrogen bonds direct the 2.3.1. GNAT family
methyl group(s) within monomethyl- and dimethyl-lysine substrates For the GNAT family, initial structural and kinetic data revealed an
away from the Fe(II) center. It has also been suggested that Ser-288 in ordered sequential bi–bi kinetic mechanism for acetyl-transfer [71]. In
JMJD2A, which is often substituted by alanine in other JHDM family this mechanism, acetyl-CoA and then the lysine substrate bind to form
members (e.g. Ala-291 in JMJD2D/KDM4D [10]), modulates the a ternary complex. An active-site glutamate (Glu-173 in yGcn5/
specificity of certain demethylases for trimethyl- versus dimethyl- ScKAT2) general base activates the ɛ-amine of lysine for direct
or monomethyl-lysine substrates [60]. nucleophilic attack of the carbonyl carbon within acetyl-CoA, forming
With the discovery of several histone lysine demethylase classes, a tetrahedral intermediate [72]. This intermediate then collapses to
future work is needed to address several issues. Where rates have acetyl-lysine product and CoA (Fig. 4).
been measured, the JHDM family catalyzes demethylation with kcat
values of ∼0.013 min− 1 [60], whereas rates of LSD1/KDM1 catalyzed 2.3.2. MYST family
demethylation are N200-fold faster at ∼ 3.1 min− 1 [64]. In addition, Initial structural and kinetic data on the MYST family member,
other members of the Fe(II)/2-oxoglutarate-dependent dioxygenase yEsa1/ScKAT5 [10], by Yan et al. in 2000 revealed an active-site
superfamily possess kcat values two to four orders of magnitude glutamate (Glu-338) that structurally aligned with the corresponding
greater than the JHDM family [65–68]. These slow demethylation glutamate residue in yGcn5/ScKAT2, suggesting the MYST family
rates suggest that JHDM demethylases may work in unidentified utilizes a similar sequential mechanism of acetyl-transfer to that of the
protein complexes or that other substrates exist in vivo. Alterna- GNAT family [73]. However, in 2002 Yan et al. published a structure of
tively, only a very small fraction of the in vitro characterized JHDM yEsa1/ScKAT5 containing a proposed acetyl-cysteine (Cys-304) inter-
enzymes may be active under these conditions assayed. Another mediate, which along with some kinetic evidence suggested a ping-
unresolved issue is that both LSD1/KDM1 and the JHDM family pong mechanism of catalysis [74]. Surprisingly, a more recent report
produce hydrogen peroxide and formaldehyde. As these compounds revealed that mutation of Cys-304 to alanine or serine decreased
are known to damage DNA within the cell nucleus, pathways must acetyltransferase activity (kcat) by only two and 10-fold, respectively
exist to protect DNA. Towards this end, it is hypothesized that [75]. This report also found no evidence of a competent acetyl-enzyme
formaldehyde can be recycled to become the reactive methyl in intermediate during the acetyl-transfer reaction. Steady-state kinetic
AdoMet [69]. Furthermore, it is possible that the human genome analyses of yEsa1/ScKAT5 indicated a sequential mechanism in which
encodes yet unidentified classes of histone demethylases, which acetyl-CoA binds prior to the lysine substrate. In addition, pH rate
work in concert with the existing LSD1/KDM1 and JHDM demethy- analysis and mutagenesis indicated Glu-338 is the general base
lase families. responsible for activating the ɛ-amine of lysine, as was originally

Fig. 4. Proposed chemical mechanism of histone acetyltransferases.


50 B.C. Smith, J.M. Denu / Biochimica et Biophysica Acta 1789 (2009) 45–57

suggested. Therefore, the MYST family utilizes an ordered sequential active-site contains two His–Asp dyads that are thought to work as a
bi–bi mechanism similar to the GNAT family. general acid/base catalytic pair [83–85]. Kinetic, structural, and
computational data suggest a mechanism in which first an active-
2.3.3. p300/CBP family site metal ion and Tyr-306 (HDAC8 numbering) polarize the carbonyl
Investigations of the p300/CBP family have led to contradictory by coordinating to the acetyl oxygen. For discussion purposes we will
findings with regard to their kinetic mechanism and structural relation assume that this active-site metal is Zn2+, but there is significant
to the GNAT and MYST families. In initial studies, a bi-substrate analog of evidence that this metal may instead be Fe2+ [86]. The Zn2+ is
acetyl-CoA and lysine was shown to be a potent inhibitor of p300/ coordinated by Asp-178, His-180, and Asp-267 [83–85]. Structural
hKAT3B [10] (IC50 ∼ 500 nM) suggesting a sequential mechanism similar studies have suggested that the Zn2+ also coordinates and activates the
to the other HAT families [76]. However, steady-state bi-substrate kinetic water molecule for attack of the acetyl carbonyl carbon, but
experiments with p300 revealed a parallel line pattern consistent with theoretical calculations indicate that this water molecule is bound
ping-pong mechanism, which was further supported by dead-end and activated by Asp-178, His-142, and His-143 [87]. In the first
inhibitor studies [77]. Recently, Liu et al. reported a crystal structure of chemical step, both His-142 [83–85,88] and His-143 [87] have been
the p300/hKAT3B HAT domain in complex with the bi-substrate analog implicated as the general base that activates the water molecule for
(see above) [78], revealing that the p300/CBP structure is distantly nucleophilic attack of the acetyl carbonyl carbon. Water attack results
related to that of the GNAT and MYST families. From this structure and in a tetrahedral intermediate that is stabilized by Zn2+ and Tyr-306
accompanying biochemical data, the authors propose that the p300/CBP [83–85]. Interestingly, in class IIa HDACs this tyrosine residue is
family utilizes a Theorell–Chance mechanism (a specialized case of an replaced by histidine, which causes a significant decrease in catalytic
ordered sequential mechanism in which the ternary complex possesses activity compared to other HDAC classes [89]. Subsequently, the
a very short lifetime), which would explain the previously observed tetrahedral intermediate collapses to form acetate and lysine
parallel line pattern. Therefore, all characterized HAT families follow an products, with His-143 acting as a general acid to protonate the ɛ-
ordered sequential bi–bi kinetic mechanism where differences between amine leaving group [83,87,88]. Theoretical studies indicate that
families may affect substrate specificity but not the overall mechanism simultaneous protonation of His-142 and His-143 within HDAC8 is
of catalysis. However, the catalytic mechanisms of the acetyltransferases unlikely; as a result, if His-142 is the general base in the first chemical
Hat1/KAT1 and Rtt109/KAT11 [10] remain unclear. Recent evidence step then His-143 must be deprotonated initially and then transfer the
suggests that Rtt109/KAT11 is a structural p300/CBP homolog [166]. proton from His-142 to His-143 [88]. However, if His-143 is the general
base, then the requirement of this proton transfer step is eliminated
2.4. Histone deacetylases [87]. Ionizations observed from pH profiles of classes I and IIb are
consistent with this proposed chemical mechanism [90], but future
Histone deacetylases (HDACs) catalyze the removal of acetyl studies are needed to unequivocally assign the individual pKa values to
groups from the ɛ-amine of acetyl-lysine residues within histones. specific active-site residues. In addition to the Zn2+ binding site,
Phylogenetic analyses subdivide HDACs into four classes. Classes I, II, structural studies of HDAC8 have identified a second monovalent
and IV utilize an active-site metal dependent catalytic mechanism, metal binding site (K+ or Na+) that is important for structural stability
and class I HDACs include human HDAC1–3 and HDAC8, class II and may play a role in catalysis, as the metal ion is in close proximity
include human HDAC4–7 and HDAC9–10, and class IV HDACs include (7.0 Å) to the active-site zinc [84]. However, future work is needed to
human HDAC11 and are homologous to both classes I and II [79]. Class determine the precise function of this monovalent metal binding site.
III HDACs (or sirtuins) utilize a distinct nicotinamide adenine Within class I/II/IV HDACs, future work will be necessary to
dinucleotide (NAD+) dependent catalytic mechanism and are con- determine the similarities and differences between individual HDAC
served from bacteria to humans with five yeast homologs (ySir2 and enzymes in terms of their catalytic mechanism and substrate
Hst1–4) and seven human homologs (Sirt1–7) [80,81]. specificity. Current knowledge is limited as the majority of in vitro
mechanistic studies have been performed with HDAC8 due to the
2.4.1. Class I/II/IV histone deacetylases difficultly in purifying other HDACs in an enzymatically active form
Class I/II/IV HDACs have homologous active-sites and are thought [91]. This difficulty likely stems from the requirement of other
to proceed through similar catalytic mechanisms (Fig. 5) [82]. The accessory proteins for full HDAC activity.

Fig. 5. Proposed chemical mechanism of class I/II/IV HDACs.


B.C. Smith, J.M. Denu / Biochimica et Biophysica Acta 1789 (2009) 45–57 51

2.4.2. Sirtuin (class III) histone deacetylases acetyl-lysine attack producing a distinct oxacarbenium intermediate
In stark contrast to class I/II/IV HDACs, sirtuins require NAD+ in a [96], while other studies suggest an SN2 mechanism where nicotina-
1:1 stoichiometry with acetyl-lysine for deacetylation, producing mide-ribosyl bond cleavage and acetyl-lysine attack occur in the same
nicotinamide and 2′-O-acetyl-ADP-ribose (OAADPr) [92,93] in addi- step [102–104]. Consistent with (but not proof of) an SN2 mechanism,
tion to deacetylated lysine as products (Fig. 6). Kinetic studies indicate electron-withdrawing fluorinated acetyl-lysine analog substrates
that sirtuins catalyze an ordered sequential bi–ter catalytic mechan- displayed significantly slower single-turnover rates of nicotinamide
ism in which acetyl-lysine substrate binds first followed by NAD+ to formation and steady-state deacetylation rates (kcat) suggesting that
form a ternary complex [94]. Structural evidence suggests that acetyl- nicotinamide-ribosyl bond cleavage step is directly tied to the
lysine-binding positions the nicotinamide ring of NAD+ within a nucleophilicity of the acetyl oxygen [104]. The actual NAD+ cleavage
highly conserved enzyme pocket where the carboxamide of NAD+ mechanism may lie somewhere on the continuum between these two
hydrogen bonds to a conserved aspartate residue (Asp-118 in Hst2) extremes. In the subsequent chemical step, a conserved histidine (His-
[95,96]. Once the ternary complex is formed, the initial chemical step 135 in Hst2) general base activates the 2′-hydroxyl for attack of the O-
involves cleavage of the nicotinamide-ribosyl bond of NAD+ and the alkylamidate imidate carbon to generate the 1′,2′-cyclic intermediate.
attack of the acetyl oxygen to form the α-1′-O-alkylamidate Mutagenesis of this histidine to alanine disables deacetylation but not
intermediate and the first product, nicotinamide. Several lines of nicotinamide exchange [97], indicating acetyl-transfer and cleavage of
evidence support the formation of this O-alkylamidate intermediate. the nicotinamide-ribosyl bond are two distinct chemical steps. This
First, in the presence of both acetyl-lysine peptide and NAD+, two-step mechanism is further supported by rapid quenching studies
exogenously added 14C-nicotinamide is rapidly incorporated to form with Hst2 that demonstrated the nicotinamide-ribosyl bond is cleaved
14
C-NAD+ in a process termed the nicotinamide exchange reaction at a rate of 8 s− 1, while the rate of subsequent acetyl-transfer to the 2′-
[97–99]. The acetyl-lysine substrate is absolutely required for efficient hydroxyl occurred at a rate of 2 s− 1 [94]. Once the 1′,2′-cyclic
cleavage of the nicotinamide-ribosyl bond in both the nicotinamide intermediate is formed, multiple steps involving water attack on the
exchange and deacetylation reactions. Replacement of NAD+ with 2′- 1′,2′-cyclic intermediate and general acid catalyzed elimination of the
deoxy-2′-fluoro-NAD+ halted the deacetylation reaction but not the ɛ-amine of lysine yields the final two products 2′-OAADPr and
nicotinamide exchange reaction [97] indicating that the first chemical deacetylated lysine. Although 2′-OAADPr is the immediate enzymatic
step occurs at the 1′-carbon of the nicotinamide ribose. Second, product, non-enzymatic interconversion between 2′-OAADPr and 3′-
oxygen labeling studies revealed transfer of an 18O label from the OAADPr is observed in solution [92,93].
acetyl oxygen to the 1′-hydroxyl in OAADPr consistent with the While much progress has been made to elucidate the sirtuin
formation of the α-1′-O-alkylamidate intermediate [93,100]. Third, catalytic mechanism, further studies are needed to determine the
replacement of the acetyl oxygen in an acetyl-lysine histone peptide mode of initial acetyl-lysine attack. Utilizing kinetic isotope effects,
with sulfur yielded a slow sirtuin substrate that permitted mass bacterial toxin ADPr-transfer reaction mechanisms have been deter-
spectral detection of a species consistent with an α-1′-S-alkylamidate mined to proceed through oxacarbenium/SN1 mechanisms of NAD+
intermediate [101]. cleavage [105–110]. Similar kinetic isotope effects performed with
Although there is significant evidence for the O-alkylamidate sirtuins would be valuable in determining the transitions state(s)
intermediate, the mechanism of its formation is a matter of debate. In leading up to O-alkylamidate formation. It is also intriguing that NAD+
particular, some studies suggest an SN1 mechanism in which is utilized in a deacetylation reaction that is already thermodynami-
nicotinamide-ribosyl bond cleavage occurs in a distinct step from cally favored. One possible reason is that utilization of NAD+ allows the

Fig. 6. Proposed chemical mechanism of class III HDACs (or sirtuins).


52 B.C. Smith, J.M. Denu / Biochimica et Biophysica Acta 1789 (2009) 45–57

coupling of a metabolic parameter (NAD+ levels) to gene silencing. coronary heart disease [6]. Protein arginine deiminases have been
There is also significant evidence that the product OAADPr possesses linked to multiple sclerosis [135] and rheumatoid arthritis [4]. Given
biologically important functions [111–114]. these essential roles in both normal cell function as well as
Several reports have suggested that in addition to deacetylase pathogenesis, it is important to understand the precise chemical
activity, some sirtuins possess NAD+-dependent ADP-ribosyltransfer- mechanisms catalyzed by these enzymes.
ase activity [81,115–119]. However, these studies have yet to
demonstrate that sirtuins are capable of multiple rounds of ADP- 3.1. Protein arginine methyltransferases
ribosylation, putting into question the biological relevance of this
activity. Additionally, only limited evidence suggests that this is Protein arginine methyltransferases (PRMTs) catalyze methyl
indeed an ADP-ribosylation reaction as the majority of evidence relies transfer from S-adenosyl-L-methionine (AdoMet) to the guanidinium
on the detection of a 32P label transferred from 32P-NAD+ to the side chain of arginine residues resulting in methylated arginine and S-
protein. To date, the specific amino acid site modified and the nature adenosyl-L-homocysteine (AdoHcy). PRMTs are conserved from yeast
of the linkage to the ADP-ribose portion of NAD+ remain unclear. A to humans, but are not found in bacteria. Mammals have at least nine
recent study suggests two main mechanisms for the observed ADP- members of this family (PRMT1–9), and are separated into two main
ribosylation, both of which require NAD+ and an acetyl-lysine types. Whereas both types catalyze the formation of monomethylar-
substrate [120]. In the first mechanism, acetyl-lysine and NAD+ react ginine (MMA), type I PRMTs (e.g. PRMT1, PRMT3, CARM1/PRMT4,
within the sirtuin active-site to form the O-alkylamidate intermediate. PRMT6, and PRMT8) continue to form asymmetric N,N′-dimethyl-
It has been shown previously that this intermediate can react with arginine (ADMA) and type II PRMTs (e.g. PRMT5, PRMT7, and FBXO11/
exogenous alcohols suggesting that amino acid side chains may react PRMT9) continue of form symmetric N,N-dimethylarginine (SDMA).
with this intermediate to yield protein ADP-ribosylation [100]. In the PRMT2 was identified by sequence homology, but no methyltransfer-
second mechanism, sirtuins react with acetyl-lysine and NAD+ to form ase activity is known.
OAADPr, which then reacts non-enzymatically with a protein resulting Structural evidence suggests that PRMT enzymes catalyze an
in ADP-ribosylation. In cases where rates of both activities have been ordered sequential bi–bi kinetic mechanism in which AdoMet binds
determined, deacetylase activity was three to five orders of magnitude prior to the arginine substrate [136]. A conserved glutamate and
greater than ADP-ribosylation activity [120,121]. The large difference arginine residue (Glu-100 and Arg-54 in PRMT1) interacts with the
in catalytic efficiency as well as the observed non-enzymatic reaction two-ribose hydroxyls and carboxylate of AdoMet, respectively. Similar
of OAADPr put into question the physiological significance of ADP- to histone lysine methyltransferases, the methylsulfonium group of
ribosylation by sirtuins that also possess deacetylase activity. Inter- AdoMet is positioned at the base of a channel in which the arginine
estingly, human Sirt4 and Sirt6 have been suggested to possess ADP- substrate binds. Two invariant glutamate residues (Glu-144 and Glu-
ribosyltransferase activity [117,118]. No deacetylase activity has been 153 in PRMT1) hydrogen bond to and position the guanidinium side
observed for Sirt4, while there is only one report of Sirt6 displaying chain of the arginine substrate. These hydrogen bonds are hypothe-
deacetylase activity [122]. Future studies are needed to determine if sized to localize the positive charge to one terminal nitrogen of the
these sirtuins can catalyze bona fide ADP-ribosylation. guanidinium side chain leaving a lone pair on the other nitrogen to
attack the methylsulfonium of AdoMet [137]. Both of these glutamate
2.5. Other histone lysine post-translational modifications residues are critical for activity as mutation of either to glutamine
reduces methyltransferase activity by ≥3000-fold and mutation to
While methylation and acetylation remain the predominant aspartate reduces activity by ≥10-fold [138]. Similar to histone lysine
modifications present on histone lysine residues, other modifications methyltransferases, methyl transfer proceeds through an SN2
have been reported, including ADP-ribosylation, ubiquitination, mechanism (Fig. 1). The proton elimination step after methyl transfer
sumoylation, biotinylation, propionylation, and butyrylation [123– is hypothesized to occur through a His–Asp proton relay system [137].
127]. Enzymes that are responsible for the cleavage of some of these As mentioned above, type I PRMTs catalyze the formation of ADMA
modifications are also known, including ubiquitin hydrolases and poly whereas type II PRMTs catalyze the formation of SDMA. In the
(ADP-ribose)glycohydrolases. In addition, in vitro evidence indicates formation of ADMA or SDMA, available in vitro data is most consistent
that class III HDACs are capable of depropionylation and debutyryla- with a partially processive mechanism in which AdoHcy release and
tion in addition to deacetylation [104,128,129]. Because modifications AdoMet binding occurs on the same time scale as release of MMA
on histone lysine residues are mutually exclusive (except perhaps containing products. In other words, the dimethyl-arginines are
monomethylation and acylation), it will be critical to determine under produced processively without release of MMA between methylation
what conditions each of these modifications occur and what steps but not in an obligate fashion as MMA products are also released
physiological response each initiates. in vitro [139,140]. However, PRMT substrates isolated in vivo are
almost completely dimethylated [141–144]. The available PRMT
3. Arginine modifying enzymes crystal structures (rat PRMT1, rat PRMT3, yeast RMT1/hmt1, and
CARM1/PRMT4) provide some insight into the processive formation of
Arginine residues within histones are subject to methylation and dimethyl-arginines. All structurally characterized PRMTs form a
citrullination (deimination) of their guanidinium side chains, cata- ringlike dimer, and dimer formation of PRMTs is thought necessary
lyzed by protein arginine methyltransferases and protein arginine for methyltransferase activity [137,138,145,146]. In addition, PRMT7
deiminases, respectively. Regulation of histone arginine methylation and PRMT8 contain two conserved catalytic core regions each
has been linked to a variety of important cellular processes including containing an AdoMet binding motif essentially constituting a dimer
transcriptional regulation, translation, and DNA repair [130]. Although [32,147]. While dimerization influences AdoMet binding [32], another
less characterized than histone lysine methylation, histone arginine potential function is to allow processive production of ADMA or
methylation can be correlated with gene repression or activation SDMA. Therefore, it is possible that this dimer structure exhibits
depending on the site and degree of methylation [131]. The biological processivity by allowing the product of the first methylation reaction
effects of histone citrullination are less clear, but citrullination has to enter the active-site of the second molecule of the dimer without
been suggested to be involved in terminal differentiation and releasing MMA. In support of this hypothesis, when the site of PRMT
apoptosis [132]. In addition to the important roles in normal cell dimerization is deleted [138], replaced with alanine residues [146], or
function, protein arginine methyltransferases are overexpressed in otherwise mutated [148], these PRMTs lack methyltransferase activity.
human prostate carcinoma [133,134] and implicated in implicated in Additionally, PRMT1 can also form oligomers in vitro and in vivo
B.C. Smith, J.M. Denu / Biochimica et Biophysica Acta 1789 (2009) 45–57 53

[138,149], which could contribute to processive mechanisms on a with respect to the arginine substrate and is comparable to JHDM
larger scale. Since the N-terminal tails of histones contain multiple family lysine demethylase activity. The requirement of an immuno-
arginines, a PRMT1 oligomer could methylate multiple arginines precipitated peptide prior to mass spectral analysis suggests a very
simultaneously, a different form of processivity. slow rate of catalysis and that only a small percentage of the substrate
was demethylated in the reported assays. In addition, these
3.2. Histone arginine demethylases immunoprecipitated peptides revealed significant oxidation of two
lysine residues, suggesting the oxidation activity of JMJD6 is not
Recently, Chang et al. reported that JMJD6, a homolog of the JHDM specific for methylated arginine residues. The physiological signifi-
family of histone lysine demethylases, possesses histone arginine cance of this lysine oxidation will also need to be addressed.
demethylase activity in lieu of lysine demethylase activity [150]. In
particular, the authors observed a reduction in dimethylated Arg-2 of 3.3. Protein arginine deiminases
histone H3 (H3R2) and dimethylated Arg-3 of histone H4 (H4R3) with
concurrent formation of 3H-formaldehyde in the presence of JMJD6. Protein arginine deiminases (PADs) catalyze the hydrolysis of
However, no change in dimethylated Arg-17 (H3R17) and Arg-26 of H3 guanidinium side chains of histone arginine residues to form citrulline
(H3R26) was observed suggesting the effects of JMJD6 were specific. and ammonia. PADs belong to a larger group of guanidinium-
Consistent with the in vitro dimethylation results, HeLa cells over- modifying enzymes termed the amidinotransferase superfamily and
expressing JMJD6 also revealed a reduction in the levels of thus are related to non-peptidyl arginine deiminase (ADI) and
dimethylated H3R2 and H4R2 but not dimethylated H3R17 or dimethyl-arginine dimethylaminohydrolase (DDAH). To date, five
H3R26. Demethylase activity was dependent on the presence of Fe human PAD homologs have been identified designated as PAD1-4
(II) and 2-oxoglutarate, and the three residues predicted to bind Fe(II). and PAD6.
In addition, mass spectrometry of a dimethyl-arginine containing Within PAD enzymes, the guanidinium side chain is held by
peptide incubated with JMJD6 and then immunoprecipitated with a hydrogen-bonding interactions with two conserved aspartate resi-
monomethylarginine specific antibody suggested the loss of one dues (Asp-350 and Asp-473 in PAD4) (Fig. 7). In the first chemical
methyl group from the dimethylated substrate. Given the predicted step, a conserved thiolate (Cys-645) whose charge is stabilized
conservation of structural elements and key catalytic residues, JMJD6 through an ion pair with His-471 undergoes nucleophilic attack at
is hypothesized to possess a similar demethylation mechanism to the guanidinium carbon to form a tetrahedral intermediate. This
JHDM lysine demethylases (Fig. 3). tetrahedral intermediate has been observed in a crystal structure of
While this first report of arginine demethylase activity is exciting, ADI co-crystallized with arginine [151]. Haloacetamine mechanism-
several issues should be resolved before bone fide arginine demethy- based inhibitors that covalently modify Cys-645 corroborate
lase activity is assigned. Of primary importance is establishing a nucleophilic attack by Cys-645 [152,153]. The pH profile of PAD4
catalytic rate of JMJD6 arginine demethylation that exhibits saturation reveals two key ionizable groups of pKa 7.3 and 8.2 that must be

Fig. 7. Proposed chemical mechanism of PAD enzymes. This figure was adapted from Thompson and Fast [159].
54 B.C. Smith, J.M. Denu / Biochimica et Biophysica Acta 1789 (2009) 45–57

deprotonated and protonated for activity, respectively. These ‘decitrullinase’ exists for the conversion of non-peptidyl citrulline to
ionizations are consistent with a reverse protonation mechanism arginine by arginosuccinate synthase and argininosuccinate lyase in
in which Cys-645 is deprotonated and His-471 is protonated prior to the urea cycle.
substrate binding [154]. After formation of this tetrahedral inter-
mediate, general acid catalyzed elimination of ammonia by His-471 4. Conclusions and perspectives
results in the S-alkylthiouronium intermediate. A structure of the S-
alkylthiouronium intermediate bound to ADI demonstrates that the While the last decade has seen significant progress in the
hydrogen-bonding network between the two aspartate residues and diverse and complex chemical mechanisms catalyzed by histone
reaction intermediates is maintained throughout the reaction lysine and arginine modifying enzymes, there is still much to be
[151,155]. Once the S-alkylthiouronium intermediate is formed, an discovered. For example, many of these enzymes form complexes
ordered water molecule (observed in the ADI/S-alkylthiouronium with other proteins containing histone-binding domains specific for
structure) attacks this intermediate passing through a tetrahedral a particular post-translational modification or contain these
intermediate to then generate citrulline as the second product. domains within their primary sequence. Therefore, determining
While the mechanism of water activation is a matter of debate, how these binding domains affect catalytic activity, substrate
available data is most consistent with His-471 acting as the general specificity, and enzyme targeting will be essential. The combinator-
base [154,156,157]. However, mechanisms in which the arginine ial action of these binding and enzymatic domains may enact part
substrate [155] or the ammonia product [157] acts as the general of the “histone code”, which suggests histone modifications act in a
base have also been proposed. While most mechanistic studies on combinatorial and sequence-dependent fashion to yield specific
PADs have been performed with PAD4, it will be of interest to downstream events [163,164].
determine if other PADs catalyze similar mechanisms or possess The involvement of these histone lysine and arginine modifying
distinct in vivo substrates. enzymes in human disease indicates that study of these enzymes is
In addition, PAD enzymes are activated by calcium, with PAD4 useful therapeutically. With this in mind, it will be interesting to
exhibiting positive cooperativity with calcium concentrations in the follow small-molecules targeting these enzymes currently in clinical
mid to high micromolar range [157]. The structures of calcium-free trials to determine if these compounds will be efficacious or if more
wild-type PAD4 and a calcium-bound PAD4 C645A mutant with specific compounds are desired. In the latter case, further elucidation
and without arginine substrate reveal that calcium binding induces of differences in mechanism and substrate specificity between these
the formation of the active-site cleft and thus is essential for enzymes will be important. Furthermore, it is possible that other
catalysis [156]. This suggests a potential connection between PAD homologous enzymes or enzymes that catalyze new chemistries on
activity and apoptosis when tight control of calcium concentrations histones (e.g. conversion of citrulline to arginine) have yet to be
is lost [158]. discovered. Even though many questions still remain, it is clear that
It has been hypothesized that PADs might serve an analogous the post-translational modification of histone lysine and arginine
function to protein arginine demethylases in the removal of methyl residues is a critical signal-transduction pathway reminiscent of other
groups (demethylimination) within histones by converting methy- well-established cascades [165].
lated arginines to citrulline. This hypothesis has been strengthened by
the observation that the sites of deimination by PAD4 overlap with the Acknowledgments
sites of arginine methylation within histones (see [159] and references
therein). However, in vivo and in vitro studies into PAD catalyzed This work was supported by National Institutes of Health grant
demethylimination are contradictory. In vivo evidence reveals an GM065386 (to J.M.D.) and by National Institutes of Health Biotechnol-
increase in histone citrulline levels that correlates with a decrease in ogy Training Grant NIH 5 T32 GM08349 (to B.C.S.). We thank
methylated arginine levels [160,161]. In contrast, in vitro studies using Christopher Berndsen, William Hallows, and Kelly Hoadley for
synthetic substrates containing MMA, ADMA, or SDMA are processed contributive discussions.
by PAD2, PAD3, and PAD4 with rates two to four orders of magnitude
slower than substrates containing unmodified arginine [157,159,162].
Therefore, PAD enzymes are relatively inefficient demethyliminases in References
vitro. However, it is possible that accessory proteins or post-
[1] R.D. Kornberg, Y. Lorch, Twenty-five years of the nucleosome, fundamental
translational modifications enhance PAD activity toward methylated particle of the eukaryote chromosome, Cell 98 (1999) 285–294.
arginine substrates. [2] K. Luger, J.C. Hansen, Nucleosome and chromatin fiber dynamics, Curr. Opin.
Struct. Biol. 15 (2005) 188–196.
Consistent with a lack of demethyliminase activity, structural and
[3] C.L. Peterson, M.A. Laniel, Histones and histone modifications, Curr. Biol. 14
kinetic data obtained with PAD4 and the related enzyme DDAH, which (2004) R546–R551.
is highly selective for methylated arginine substrates, lend insight into [4] A. Suzuki, R. Yamada, K. Yamamoto, Citrullination by peptidylarginine deiminase
the ability of PAD4 to discriminate between substrates differing by one in rheumatoid arthritis, Ann. N. Y. Acad. Sci. 1108 (2007) 323–339.
[5] M.A. Glozak, E. Seto, Histone deacetylases and cancer, Oncogene 26 (2007)
methyl group. A structural overlay of the PAD4 and DDAH active-sites 5420–5432.
reveal one notable difference in the guanidinium-binding pocket [6] X. Chen, F. Niroomand, Z. Liu, A. Zankl, H.A. Katus, L. Jahn, C.P. Tiefenbacher,
[159]. PAD4 hydrogen bonds to the two terminal nitrogens via Asp- Expression of nitric oxide related enzymes in coronary heart disease, Basic Res.
Cardiol. 101 (2006) 346–353.
473, whereas DDAH contains a lysine in this position forming a part of [7] V.D. Longo, B.K. Kennedy, Sirtuins in aging and age-related disease, Cell 126
a pocket that accommodates N-methyl groups. In PAD4, this Lys-to- (2006) 257–268.
Asp substitution along with changes in the positioning of Val-469, [8] J.C. Milne, J.M. Denu, The Sirtuin family: therapeutic targets to treat diseases of
aging, Curr. Opin. Chem. Biol. (2008).
Glu-474, and Asn-588 are thought to help prevent N-methyl binding [9] T.S. Anekonda, P.H. Reddy, Neuronal protection by sirtuins in Alzheimer's disease,
within the active-site, resulting in selectivity of PAD4 for unmethy- J. Neurochem. 96 (2006) 305–313.
lated arginine substrates. [10] C.D. Allis, S.L. Berger, J. Cote, S. Dent, T. Jenuwien, T. Kouzarides, L. Pillus, D.
Reinberg, Y. Shi, R. Shiekhattar, A. Shilatifard, J. Workman, Y. Zhang, New
In vivo evidence that histone H3 and H4 citrulline levels rise and
nomenclature for chromatin-modifying enzymes, Cell 131 (2007) 633–636.
fall suggests that citrullination may be a reversible modification. [11] J.C. Rice, C.D. Allis, Histone methylation versus histone acetylation: new insights
Variations in citrulline levels could be due to histone tail clipping, into epigenetic regulation, Curr. Opin. Cell Biol. 13 (2001) 263–273.
[12] R. Schneider, A.J. Bannister, T. Kouzarides, Unsafe SETs: histone lysine
epitope occlusion, or nucleosome displacement. Alternatively, these
methyltransferases and cancer, Trends Biochem. Sci. 27 (2002) 396–402.
variations could be due to an undiscovered aminotransferase that [13] O. Van Beekum, E. Kalkhoven, Aberrant forms of histone acetyltransferases in
converts citrulline to arginine within histones. Precedence for such a human disease, Subcell. Biochem. 41 (2007) 233–262.
B.C. Smith, J.M. Denu / Biochimica et Biophysica Acta 1789 (2009) 45–57 55

[14] P.A. Marks, R. Breslow, Dimethyl sulfoxide to vorinostat: development of this [45] J. Min, Q. Feng, Z. Li, Y. Zhang, R.M. Xu, Structure of the catalytic domain of human
histone deacetylase inhibitor as an anticancer drug, Nat. Biotechnol. 25 (2007) DOT1L, a non-SET domain nucleosomal histone methyltransferase, Cell 112
84–90. (2003) 711–723.
[15] I. Letunic, R.R. Copley, S. Schmidt, F.D. Ciccarelli, T. Doerks, J. Schultz, C.P. Ponting, [46] K. Sawada, Z. Yang, J.R. Horton, R.E. Collins, X. Zhang, X. Cheng, Structure of the
P. Bork, SMART 4.0: towards genomic data integration, Nucleic Acids Res. 32 conserved core of the yeast Dot1p, a nucleosomal histone H3 lysine 79
(2004) D142–144. methyltransferase, J. Biol. Chem. 279 (2004) 43296–43306.
[16] H.G. Chin, D. Patnaik, P.O. Esteve, S.E. Jacobsen, S. Pradhan, Catalytic properties [47] H.L. Schubert, R.M. Blumenthal, X. Cheng, Many paths to methyltransfer:
and kinetic mechanism of human recombinant Lys-9 histone H3 methyltransfer- a chronicle of convergence, Trends Biochem. Sci. 28 (2003) 329–335.
ase SUV39H1: participation of the chromodomain in enzymatic catalysis, [48] W.K. Paik, S. Kim, Enzymatic demethylation of calf thymus histones, Biochem.
Biochemistry 45 (2006) 3272–3284. Biophys. Res. Commun. 51 (1973) 781–788.
[17] D. Patnaik, H.G. Chin, P.O. Esteve, J. Benner, S.E. Jacobsen, S. Pradhan, Substrate [49] F. Forneris, C. Binda, A. Adamo, E. Battaglioli, A. Mattevi, Structural basis of LSD1-
specificity and kinetic mechanism of mammalian G9a histone H3 methyltrans- CoREST selectivity in histone H3 recognition, J. Biol. Chem. 282 (2007)
ferase, J. Biol. Chem. 279 (2004) 53248–53258. 20070–20074.
[18] H.B. Guo, H. Guo, Mechanism of histone methylation catalyzed by protein lysine [50] F. Forneris, C. Binda, M.A. Vanoni, A. Mattevi, E. Battaglioli, Histone demethyla-
methyltransferase SET7/9 and origin of product specificity, Proc. Natl. Acad. Sci. tion catalysed by LSD1 is a flavin-dependent oxidative process, FEBS Lett. 579
U. S. A. 104 (2007) 8797–8802. (2005) 2203–2207.
[19] Z.S. Derewenda, L. Lee, U. Derewenda, The occurrence of C–H…O hydrogen bonds [51] F. Forneris, C. Binda, A. Dall'Aglio, M.W. Fraaije, E. Battaglioli, A. Mattevi, A highly
in proteins, J. Mol. Biol. 252 (1995) 248–262. specific mechanism of histone H3-K4 recognition by histone demethylase LSD1,
[20] S. Scheiner, T. Kar, Y. Gu, Strength of the Calpha H…O hydrogen bond of amino J. Biol. Chem. 281 (2006) 35289–35295.
acid residues, J. Biol. Chem. 276 (2001) 9832–9837. [52] S. Ghisla, Fluorescence and optical characteristics of reduced flavins and
[21] H. Park, J. Yoon, C. Seok, Strength of Calpha–H…OfC hydrogen bonds in flavoproteins, Methods Enzymol. 66 (1980) 360–373.
transmembrane proteins, J. Phys. Chem. B 112 (2008) 1041–1048. [53] J.C. Culhane, P.A. Cole, LSD1 and the chemistry of histone demethylation, Curr.
[22] S. Scheiner, T. Kar, Effect of solvent upon CH…O hydrogen bonds with Opin. Chem. Biol. 11 (2007) 561–568.
implications for protein folding, J. Phys. Chem. B 109 (2005) 3681–3689. [54] E.C. Ralph, M.A. Anderson, W.W. Cleland, P.F. Fitzpatrick, Mechanistic studies of
[23] P. Hu, Y. Zhang, Catalytic mechanism and product specificity of the histone lysine the flavoenzyme tryptophan 2-monooxygenase: deuterium and 15N kinetic
methyltransferase SET7/9: an ab initio QM/MM-FE study with multiple initial isotope effects on alanine oxidation by an L-amino acid oxidase, Biochemistry 45
structures, J. Am. Chem. Soc. 128 (2006) 1272–1278. (2006) 15844–15852.
[24] J.F. Couture, G. Hauk, M.J. Thompson, G.M. Blackburn, R.C. Trievel, Catalytic roles [55] R.B. Silverman, S.J. Hoffman, W.B. Catus III, A mechanism for mitochondrial
for carbon–oxygen hydrogen bonding in SET domain lysine methyltransferases, monoamine oxidase catalyzed amine oxidation, J. Am. Chem. Soc. 102 (1980)
J. Biol. Chem. 281 (2006) 19280–19287. 7126–7128.
[25] T. Kwon, J.H. Chang, E. Kwak, C.W. Lee, A. Joachimiak, Y.C. Kim, J. Lee, Y. Cho, [56] Y. Tsukada, J. Fang, H. Erdjument-Bromage, M.E. Warren, C.H. Borchers, P. Tempst,
Mechanism of histone lysine methyl transfer revealed by the structure of SET7/ Y. Zhang, Histone demethylation by a family of JmjC domain-containing proteins,
9-AdoMet, EMBO J. 22 (2003) 292–303. Nature 439 (2006) 811–816.
[26] X. Zhang, Z. Yang, S.I. Khan, J.R. Horton, H. Tamaru, E.U. Selker, X. Cheng, [57] R. Anand, R. Marmorstein, Structure and mechanism of lysine-specific demethy-
Structural basis for the product specificity of histone lysine methyltransferases, lase enzymes, J. Biol. Chem. 282 (2007) 35425–35429.
Mol. Cell. 12 (2003) 177–185. [58] R.P. Hausinger, Fe(II)/alpha-ketoglutarate-dependent hydroxylases and related
[27] S.A. Jacobs, J.M. Harp, S. Devarakonda, Y. Kim, F. Rastinejad, S. Khorasanizadeh, enzymes, Crit. Rev. Biochem. Mol. Biol. 39 (2004) 21–68.
The active site of the SET domain is constructed on a knot, Nat. Struct. Biol. 9 [59] A. Ozer, R.K. Bruick, Non-heme dioxygenases: cellular sensors and regulators
(2002) 833–838. jelly rolled into one? Nat. Chem. Biol. 3 (2007) 144–153.
[28] X. Zhang, H. Tamaru, S.I. Khan, J.R. Horton, L.J. Keefe, E.U. Selker, X. Cheng, [60] J.F. Couture, E. Collazo, P.A. Ortiz-Tello, J.S. Brunzelle, R.C. Trievel, Specificity and
Structure of the Neurospora SET domain protein DIM-5, a histone H3 lysine mechanism of JMJD2A, a trimethyllysine-specific histone demethylase, Nat.
methyltransferase, Cell 111 (2002) 117–127. Struct. Mol. Biol. 14 (2007) 689–695.
[29] L.M. Dirk, E.M. Flynn, K. Dietzel, J.F. Couture, R.C. Trievel, R.L. Houtz, Kinetic [61] S.S. Ng, K.L. Kavanagh, M.A. McDonough, D. Butler, E.S. Pilka, B.M. Lienard, J.E. Bray,
manifestation of processivity during multiple methylations catalyzed by SET P. Savitsky, O. Gileadi, F. von Delft, N.R. Rose, J. Offer, J.C. Scheinost, T. Borowski, M.
domain protein methyltransferases, Biochemistry 46 (2007) 3905–3915. Sundstrom, C.J. Schofield, U. Oppermann, Crystal structures of histone demethy-
[30] B. Xiao, C. Jing, G. Kelly, P.A. Walker, F.W. Muskett, T.A. Frenkiel, S.R. Martin, K. lase JMJD2A reveal basis for substrate specificity, Nature 448 (2007) 87–91.
Sarma, D. Reinberg, S.J. Gamblin, J.R. Wilson, Specificity and mechanism of the [62] Z. Chen, J. Zang, J. Kappler, X. Hong, F. Crawford, Q. Wang, F. Lan, C. Jiang, J.
histone methyltransferase Pr-Set7, Genes Dev. 19 (2005) 1444–1454. Whetstine, S. Dai, K. Hansen, Y. Shi, G. Zhang, Structural basis of the recognition
[31] F.H. Westheimer, Coincidences, decarboxylation, and electrostatic effects, of a methylated histone tail by JMJD2A, Proc. Natl. Acad. Sci. U. S. A. 104 (2007)
Tetrahedron 51 (1995) 3–20. 10818–10823.
[32] X. Cheng, R.E. Collins, X. Zhang, Structural and sequence motifs of protein [63] Y. Shi, J.R. Whetstine, Dynamic regulation of histone lysine methylation by
(histone) methylation enzymes, Annu. Rev. Biophys. Biomol. Struct. 34 (2005) demethylases, Mol. Cell 25 (2007) 1–14.
267–294. [64] L.M. Szewczuk, J.C. Culhane, M. Yang, A. Majumdar, H. Yu, P.A. Cole, Mechanistic
[33] C. Qian, M.M. Zhou, SET domain protein lysine methyltransferases: structure, analysis of a suicide inactivator of histone demethylase LSD1, Biochemistry 46
specificity and catalysis, Cell Mol. Life Sci. 63 (2006) 2755–2763. (2007) 6892–6902.
[34] B. Xiao, J.R. Wilson, S.J. Gamblin, SET domains and histone methylation, Curr. [65] P.K. Grzyska, M.J. Ryle, G.R. Monterosso, J. Liu, D.P. Ballou, R.P. Hausinger, Steady-
Opin. Struct. Biol. 13 (2003) 699–705. state and transient kinetic analyses of taurine/alpha-ketoglutarate dioxygenase:
[35] X. Zhang, T.C. Bruice, Histone lysine methyltransferase SET7/9: formation of a water effects of oxygen concentration, alternative sulfonates, and active-site variants
channel precedes each methyl transfer, Biochemistry 46 (2007) 14838–14844. on the FeIV-oxo intermediate, Biochemistry 44 (2005) 3845–3855.
[36] X. Zhang, T.C. Bruice, Mechanism of product specificity of AdoMet methylation [66] T.W. Roy, A.S. Bhagwat, Kinetic studies of Escherichia coli AlkB using a new
catalyzed by lysine methyltransferases: transcriptional factor p53 methylation by fluorescence-based assay for DNA demethylation, Nucleic Acids Res. 35 (2007)
histone lysine methyltransferase SET7/9, Biochemistry 47 (2008) 2743–2748. e147.
[37] P.O. Esteve, D. Patnaik, H.G. Chin, J. Benner, M.A. Teitell, S. Pradhan, Functional [67] M.D. Wolfe, D.J. Altier, A. Stubna, C.V. Popescu, E. Munck, J.D. Lipscomb, Benzoate
analysis of the N- and C-terminus of mammalian G9a histone H3 methyltrans- 1,2-dioxygenase from Pseudomonas putida: single turnover kinetics and regulation
ferase, Nucleic Acids Res. 33 (2005) 3211–3223. of a two-component Rieske dioxygenase, Biochemistry 41 (2002) 9611–9626.
[38] B. Xiao, C. Jing, J.R. Wilson, P.A. Walker, N. Vasisht, G. Kelly, S. Howell, I.A. Taylor, [68] M.J. Ryle, R. Padmakumar, R.P. Hausinger, Stopped-flow kinetic analysis of
G.M. Blackburn, S.J. Gamblin, Structure and catalytic mechanism of the human Escherichia coli taurine/alpha-ketoglutarate dioxygenase: interactions with
histone methyltransferase SET7/9, Nature 421 (2003) 652–656. alpha-ketoglutarate, taurine, and oxygen, Biochemistry 38 (1999) 15278–15286.
[39] R.C. Trievel, E.M. Flynn, R.L. Houtz, J.H. Hurley, Mechanism of multiple lysine [69] E. Tyihak, L. Trezl, B. Szende, Formaldehyde cycle and the phases of stress
methylation by the SET domain enzyme Rubisco LSMT, Nat. Struct. Biol. 10 (2003) syndrome, Ann. N. Y. Acad. Sci. 851 (1998) 259–270.
545–552. [70] S.Y. Roth, J.M. Denu, C.D. Allis, Histone acetyltransferases, Annu. Rev. Biochem. 70
[40] P. Hu, S. Wang, Y. Zhang, How do SET-domain protein lysine methyltransferases (2001) 81–120.
achieve the methylation state specificity? Revisited by ab initio QM/MM [71] K.G. Tanner, M.R. Langer, Y. Kim, J.M. Denu, Kinetic mechanism of the histone
molecular dynamics simulations, J. Am. Chem. Soc. (2008). acetyltransferase GCN5 from yeast, J. Biol. Chem. 275 (2000) 22048–22055.
[41] Q. Feng, H. Wang, H.H. Ng, H. Erdjument-Bromage, P. Tempst, K. Struhl, Y. Zhang, [72] K.G. Tanner, R.C. Trievel, M.H. Kuo, R.M. Howard, S.L. Berger, C.D. Allis, R.
Methylation of H3-lysine 79 is mediated by a new family of HMTases without a Marmorstein, J.M. Denu, Catalytic mechanism and function of invariant glutamic
SET domain, Curr. Biol. 12 (2002) 1052–1058. acid 173 from the histone acetyltransferase GCN5 transcriptional coactivator,
[42] N. Lacoste, R.T. Utley, J.M. Hunter, G.G. Poirier, J. Cote, Disruptor of telomeric J. Biol. Chem. 274 (1999) 18157–18160.
silencing-1 is a chromatin-specific histone H3 methyltransferase, J. Biol. Chem. [73] Y. Yan, N.A. Barlev, R.H. Haley, S.L. Berger, R. Marmorstein, Crystal structure of
277 (2002) 30421–30424. yeast Esa1 suggests a unified mechanism for catalysis and substrate binding by
[43] H.H. Ng, Q. Feng, H. Wang, H. Erdjument-Bromage, P. Tempst, Y. Zhang, K. Struhl, histone acetyltransferases, Mol. Cell 6 (2000) 1195–1205.
Lysine methylation within the globular domain of histone H3 by Dot1 is [74] Y. Yan, S. Harper, D.W. Speicher, R. Marmorstein, The catalytic mechanism of the
important for telomeric silencing and Sir protein association, Genes Dev. 16 ESA1 histone acetyltransferase involves a self-acetylated intermediate, Nat.
(2002) 1518–1527. Struct. Biol. 9 (2002) 862–869.
[44] F. van Leeuwen, P.R. Gafken, D.E. Gottschling, Dot1p modulates silencing in yeast [75] C.E. Berndsen, B.N. Albaugh, S. Tan, J.M. Denu, Catalytic mechanism of a MYST
by methylation of the nucleosome core, Cell 109 (2002) 745–756. family histone acetyltransferase, Biochemistry 46 (2007) 623–629.
56 B.C. Smith, J.M. Denu / Biochimica et Biophysica Acta 1789 (2009) 45–57

[76] V. Sagar, W. Zheng, P.R. Thompson, P.A. Cole, Bisubstrate analogue structure– [104] B.C. Smith, J.M. Denu, Sir2 deacetylases exhibit nucleophilic participation of
activity relationships for p300 histone acetyltransferase inhibitors, Bioorg. Med. acetyl-lysine in NAD+ cleavage, J. Am. Chem. Soc. 129 (2007) 5802–5803.
Chem. 12 (2004) 3383–3390. [105] P.J. Berti, S.R. Blanke, V.L. Schramm, Transition state structure for the hydrolysis
[77] P.R. Thompson, H. Kurooka, Y. Nakatani, P.A. Cole, Transcriptional coactivator of NAD+ catalyzed by diphtheria toxin, J. Am. Chem. Soc. 119 (1997)
protein p300. Kinetic characterization of its histone acetyltransferase activity, 12079–12088.
J. Biol. Chem. 276 (2001) 33721–33729. [106] S.L. Parikh, V.L. Schramm, Transition state structure for ADP-ribosylation of
[78] X. Liu, L. Wang, K. Zhao, P.R. Thompson, Y. Hwang, R. Marmorstein, P.A. Cole, The eukaryotic elongation factor 2 catalyzed by diphtheria toxin, Biochemistry 43
structural basis of protein acetylation by the p300/CBP transcriptional coacti- (2004) 1204–1212.
vator, Nature 451 (2008) 846–850. [107] K.A. Rising, V.L. Schramm, Transition state analysis of NAD+ hydrolysis by the
[79] I.V. Gregoretti, Y.M. Lee, H.V. Goodson, Molecular evolution of the histone cholera toxin catalytic subunit, J. Am. Chem. Soc. 119 (1997) 27–37.
deacetylase family: functional implications of phylogenetic analysis, J. Mol. Biol. [108] J. Scheuring, P.J. Berti, V.L. Schramm, Transition-state structure for the ADP-
338 (2004) 17–31. ribosylation of recombinant gia1 subunits by pertussis toxin, Biochemistry 37
[80] R.A. Frye, Phylogenetic classification of prokaryotic and eukaryotic Sir2-like (1998) 2748–2758.
proteins, Biochem. Biophys. Res. Commun. 273 (2000) 793–798. [109] J. Scheuring, V.L. Schramm, Pertussis toxin: transition state analysis for ADP-
[81] S. Imai, C.M. Armstrong, M. Kaeberlein, L. Guarente, Transcriptional silencing and ribosylation of G-protein peptide alphai3C20, Biochemistry 36 (1997)
longevity protein Sir2 is an NAD-dependent histone deacetylase, Nature 403 8215–8223.
(2000) 795–800. [110] J. Scheuring, V.L. Schramm, Kinetic isotope effect characterization of the
[82] S.C. Hodawadekar, R. Marmorstein, Chemistry of acetyl transfer by histone transition state for oxidized nicotinamide adenine dinucleotide hydrolysis by
modifying enzymes: structure, mechanism and implications for effector design, pertussis toxin, Biochemistry 36 (1997) 4526–4534.
Oncogene 26 (2007) 5528–5540. [111] M.T. Borra, F.J. O'Neill, M.D. Jackson, B. Marshall, E. Verdin, K.R. Foltz, J.M. Denu,
[83] J.R. Somoza, R.J. Skene, B.A. Katz, C. Mol, J.D. Ho, A.J. Jennings, C. Luong, A. Arvai, Conserved enzymatic production and biological effect of O-acetyl-ADP-ribose by
J.J. Buggy, E. Chi, J. Tang, B.C. Sang, E. Verner, R. Wynands, E.M. Leahy, D.R. silent information regulator 2-like NAD+-dependent deacetylases, J. Biol. Chem.
Dougan, G. Snell, M. Navre, M.W. Knuth, R.V. Swanson, D.E. McRee, L.W. Tari, 277 (2002) 12632–12641.
Structural snapshots of human HDAC8 provide insights into the class I histone [112] O. Grubisha, L.A. Rafty, C.L. Takanishi, X. Xu, L. Tong, A.L. Perraud, A.M.
deacetylases, Structure 12 (2004) 1325–1334. Scharenberg, J.M. Denu, Metabolite of SIR2 reaction modulates TRPM2 ion
[84] A. Vannini, C. Volpari, G. Filocamo, E.C. Casavola, M. Brunetti, D. Renzoni, P. channel, J. Biol. Chem. 281 (2006) 14057–14065.
Chakravarty, C. Paolini, R. De Francesco, P. Gallinari, C. Steinkuhler, S. Di Marco, [113] G. Kustatscher, M. Hothorn, C. Pugieux, K. Scheffzek, A.G. Ladurner, Splicing
Crystal structure of a eukaryotic zinc-dependent histone deacetylase, human regulates NAD metabolite binding to histone macroH2A, Nat. Struct. Mol. Biol. 12
HDAC8, complexed with a hydroxamic acid inhibitor, Proc. Natl. Acad. Sci. U. S. A. (2005) 624–625.
101 (2004) 15064–15069. [114] G.G. Liou, J.C. Tanny, R.G. Kruger, T. Walz, D. Moazed, Assembly of the SIR complex
[85] A. Vannini, C. Volpari, P. Gallinari, P. Jones, M. Mattu, A. Carfi, R. De Francesco, C. and its regulation by O-acetyl-ADP-ribose, a product of NAD-dependent histone
Steinkuhler, S. Di Marco, Substrate binding to histone deacetylases as shown by the deacetylation, Cell 121 (2005) 515–527.
crystal structure of the HDAC8–substrate complex, EMBO Rep. 8 (2007) 879–884. [115] R.A. Frye, Characterization of five human cDNAs with homology to the yeast SIR2
[86] S.L. Gantt, S.G. Gattis, C.A. Fierke, Catalytic activity and inhibition of human gene: Sir2-like proteins (sirtuins) metabolize NAD and may have protein ADP-
histone deacetylase 8 is dependent on the identity of the active site metal ion, ribosyltransferase activity, Biochem. Biophys. Res. Commun. 260 (1999) 273–279.
Biochemistry 45 (2006) 6170–6178. [116] J.A. Garcia-Salcedo, P. Gijon, D.P. Nolan, P. Tebabi, E. Pays, A chromosomal SIR2
[87] C. Corminboeuf, P. Hu, M.E. Tuckerman, Y. Zhang, Unexpected deacetylation homologue with both histone NAD-dependent ADP-ribosyltransferase and
mechanism suggested by a density functional theory QM/MM study of histone- deacetylase activities is involved in DNA repair in Trypanosoma brucei, EMBO J.
deacetylase-like protein, J. Am. Chem. Soc. 128 (2006) 4530–4531. 22 (2003) 5851–5862.
[88] K. Vanommeslaeghe, F. De Proft, S. Loverix, D. Tourwe, P. Geerlings, Theoretical [117] M.C. Haigis, R. Mostoslavsky, K.M. Haigis, K. Fahie, D.C. Christodoulou, A.J.
study revealing the functioning of a novel combination of catalytic motifs in Murphy, D.M. Valenzuela, G.D. Yancopoulos, M. Karow, G. Blander, C. Wolberger,
histone deacetylase, Bioorg. Med. Chem. 13 (2005) 3987–3992. T.A. Prolla, R. Weindruch, F.W. Alt, L. Guarente, SIRT4 inhibits glutamate
[89] W. Fischle, F. Dequiedt, M.J. Hendzel, M.G. Guenther, M.A. Lazar, W. Voelter, E. dehydrogenase and opposes the effects of calorie restriction in pancreatic beta
Verdin, Enzymatic activity associated with class II HDACs is dependent on a cells, Cell 126 (2006) 941–954.
multiprotein complex containing HDAC3 and SMRT/N-CoR, Mol. Cell 9 (2002) [118] G. Liszt, E. Ford, M. Kurtev, L. Guarente, Mouse Sir2 homolog SIRT6 is a nuclear
45–57. ADP-ribosyltransferase, J. Biol. Chem. 280 (2005) 21313–21320.
[90] B.E. Schultz, S. Misialek, J. Wu, J. Tang, M.T. Conn, R. Tahilramani, L. Wong, [119] J.C. Tanny, G.J. Dowd, J. Huang, H. Hilz, D. Moazed, An enzymatic activity in the
Kinetics and comparative reactivity of human class I and class IIb histone yeast Sir2 protein that is essential for gene silencing, Cell 99 (1999) 735–745.
deacetylases, Biochemistry 43 (2004) 11083–11091. [120] T.M. Kowieski, S. Lee, J.M. Denu, Acetylation-dependent ADP-ribosylation by
[91] N. Sengupta, E. Seto, Regulation of histone deacetylase activities, J. Cell. Biochem. Trypanosoma brucei Sir2, J. Biol. Chem. 283 (2008) 5317–5326.
93 (2004) 57–67. [121] K.G. Tanner, J. Landry, R. Sternglanz, J.M. Denu, Silent information regulator 2 family
[92] M.D. Jackson, J.M. Denu, Structural identification of 2′- and 3′-O-acetyl-ADP- of NAD-dependent histone/protein deacetylases generates a unique product, 1-O-
ribose as novel metabolites derived from the Sir2 family of beta-NAD+- acetyl-ADP-ribose, Proc. Natl. Acad. Sci. U. S. A. 97 (2000) 14178–14182.
dependent histone/protein deacetylases, J. Biol. Chem. 277 (2002) [122] E. Michishita, R.A. McCord, E. Berber, M. Kioi, H. Padilla-Nash, M. Damian, P.
18535–18544. Cheung, R. Kusumoto, T.L. Kawahara, J.C. Barrett, H.Y. Chang, V.A. Bohr, T. Ried, O.
[93] A.A. Sauve, I. Celic, J. Avalos, H. Deng, J.D. Boeke, V.L. Schramm, Chemistry of gene Gozani, K.F. Chua, SIRT6 is a histone H3 lysine 9 deacetylase that modulates
silencing: the mechanism of NAD+-dependent deacetylation reactions, Biochem- telomeric chromatin, Nature (2008).
istry 40 (2001) 15456–15463. [123] Y. Chen, R. Sprung, Y. Tang, H. Ball, B. Sangras, S.C. Kim, J.R. Falck, J. Peng, W. Gu, Y.
[94] M.T. Borra, M.R. Langer, J.T. Slama, J.M. Denu, Substrate specificity and kinetic Zhao, Lysine propionylation and butyrylation are novel post-translational
mechanism of the Sir2 family of NAD+-dependent histone/protein deacetylases, modifications in histones, Mol. Cell. Proteomics 6 (2007) 812–819.
Biochemistry 43 (2004) 9877–9887. [124] G. Gill, Something about SUMO inhibits transcription, Curr. Opin. Genet. Dev. 15
[95] J.L. Avalos, J.D. Boeke, C. Wolberger, Structural basis for the mechanism and (2005) 536–541.
regulation of Sir2 enzymes, Mol. Cell 13 (2004) 639–648. [125] P.O. Hassa, S.S. Haenni, M. Elser, M.O. Hottiger, Nuclear ADP-ribosylation
[96] K. Zhao, R. Harshaw, X. Chai, R. Marmorstein, Structural basis for nicotinamide reactions in mammalian cells: where are we today and where are we going?
cleavage and ADP-ribose transfer by NAD(+)-dependent Sir2 histone/protein Microbiol. Mol. Biol. Rev. 70 (2006) 789–829.
deacetylases, Proc. Natl. Acad. Sci. U. S. A. 101 (2004) 8563–8568. [126] A. Shilatifard, Chromatin modifications by methylation and ubiquitination:
[97] M.D. Jackson, M.T. Schmidt, N.J. Oppenheimer, J.M. Denu, Mechanism of implications in the regulation of gene expression, Annu. Rev. Biochem. 75 (2006)
nicotinamide inhibition and transglycosidation by Sir2 histone/protein deacety- 243–269.
lases, J. Biol. Chem. 278 (2003) 50985–50998. [127] Y.I. Hassan, J. Zempleni, Epigenetic regulation of chromatin structure and gene
[98] J. Landry, A. Sutton, S.T. Tafrov, R.C. Heller, J. Stebbins, L. Pillus, R. Sternglanz, The function by biotin, J. Nutr. 136 (2006) 1763–1765.
silencing protein SIR2 and its homologs are NAD-dependent protein deacety- [128] J. Garrity, J.G. Gardner, W. Hawse, C. Wolberger, J.C. Escalante-Semerena, N-lysine
lases, Proc. Natl. Acad. Sci. U. S. A. 97 (2000) 5807–5811. propionylation controls the activity of propionyl-CoA synthetase, J. Biol. Chem.
[99] A.A. Sauve, V.L. Schramm, Sir2 regulation by nicotinamide results from switching 282 (2007) 30239–30245.
between base exchange and deacetylation chemistry, Biochemistry 42 (2003) [129] B.C. Smith, J.M. Denu, Acetyl-lysine analog peptides as mechanistic probes of
9249–9256. protein deacetylases, J. Biol. Chem. 282 (2007) 37256–37265.
[100] B.C. Smith, J.M. Denu, Sir2 protein deacetylases: evidence for chemical [130] M.T. Bedford, S. Richard, Arginine methylation an emerging regulator of protein
intermediates and functions of a conserved histidine, Biochemistry 45 (2006) function, Mol. Cell 18 (2005) 263–272.
272–282. [131] J. Wysocka, C.D. Allis, S. Coonrod, Histone arginine methylation and its dynamic
[101] B.C. Smith, J.M. Denu, Mechanism-based inhibition of sir2 deacetylases by regulation, Front. Biosci. 11 (2006) 344–355.
thioacetyl-lysine peptide, Biochemistry 46 (2007) 14478–14486. [132] E.R. Vossenaar, A.J. Zendman, W.J. van Venrooij, G.J. Pruijn, PAD, a growing family
[102] K.G. Hoff, J.L. Avalos, K. Sens, C. Wolberger, Insights into the sirtuin mechanism of citrullinating enzymes: genes, features and involvement in disease, Bioessays
from ternary complexes containing NAD+ and acetylated peptide, Structure 14 25 (2003) 1106–1118.
(2006) 1231–1240. [133] H. Hong, C. Kao, M.H. Jeng, J.N. Eble, M.O. Koch, T.A. Gardner, S. Zhang, L. Li, C.X.
[103] A.N. Khan, P.N. Lewis, Use of substrate analogs and mutagenesis to study Pan, Z. Hu, G.T. MacLennan, L. Cheng, Aberrant expression of CARM1, a
substrate binding and catalysis in the Sir2 family of NAD-dependent protein transcriptional coactivator of androgen receptor, in the development of prostate
deacetylases, J. Biol. Chem. 281 (2006) 11702–11711. carcinoma and androgen-independent status, Cancer 101 (2004) 83–89.
B.C. Smith, J.M. Denu / Biochimica et Biophysica Acta 1789 (2009) 45–57 57

[134] S. Majumder, Y. Liu, O.H. Ford III, J.L. Mohler, Y.E. Whang, Involvement of arginine localization, substrate specificity, and regulation, J. Biol. Chem. 273 (1998)
methyltransferase CARM1 in androgen receptor function and prostate cancer cell 16935–16945.
viability, Prostate 66 (2006) 1292–1301. [150] B. Chang, Y. Chen, Y. Zhao, R.K. Bruick, JMJD6 is a histone arginine demethylase,
[135] F.G. Mastronardi, D.D. Wood, J. Mei, R. Raijmakers, V. Tseveleki, H.M. Dosch, L. Science 318 (2007) 444–447.
Probert, P. Casaccia-Bonnefil, M.A. Moscarello, Increased citrullination of histone [151] K. Das, G.H. Butler, V. Kwiatkowski, A.D. Clark Jr., P. Yadav, E. Arnold, Crystal
H3 in multiple sclerosis brain and animal models of demyelination: a role for structures of arginine deiminase with covalent reaction intermediates; implica-
tumor necrosis factor-induced peptidylarginine deiminase 4 translocation, tions for catalytic mechanism, Structure 12 (2004) 657–667.
J. Neurosci. 26 (2006) 11387–11396. [152] Y. Luo, K. Arita, M. Bhatia, B. Knuckley, Y.H. Lee, M.R. Stallcup, M. Sato, P.R.
[136] W.W. Yue, M. Hassler, S.M. Roe, V. Thompson-Vale, L.H. Pearl, Insights into Thompson, Inhibitors and inactivators of protein arginine deiminase 4: functional
histone code syntax from structural and biochemical studies of CARM1 and structural characterization, Biochemistry 45 (2006) 11727–11736.
methyltransferase, EMBO J. 26 (2007) 4402–4412. [153] Y. Luo, B. Knuckley, Y.H. Lee, M.R. Stallcup, P.R. Thompson, A fluoroacetamidine-
[137] X. Zhang, L. Zhou, X. Cheng, Crystal structure of the conserved core of protein based inactivator of protein arginine deiminase 4: design, synthesis, and in vitro
arginine methyltransferase PRMT3, EMBO J. 19 (2000) 3509–3519. and in vivo evaluation, J. Am. Chem. Soc. 128 (2006) 1092–1093.
[138] X. Zhang, X. Cheng, Structure of the predominant protein arginine methyl- [154] B. Knuckley, M. Bhatia, P.R. Thompson, Protein arginine deiminase 4: evidence for
transferase PRMT1 and analysis of its binding to substrate peptides, Structure 11 a reverse protonation mechanism, Biochemistry 46 (2007) 6578–6587.
(2003) 509–520. [155] A. Galkin, X. Lu, D. Dunaway-Mariano, O. Herzberg, Crystal structures
[139] A. Frankel, N. Yadav, J. Lee, T.L. Branscombe, S. Clarke, M.T. Bedford, The novel representing the Michaelis complex and the thiouronium reaction intermediate
human protein arginine N-methyltransferase PRMT6 is a nuclear enzyme of Pseudomonas aeruginosa arginine deiminase, J. Biol. Chem. 280 (2005)
displaying unique substrate specificity, J. Biol. Chem. 277 (2002) 3537–3543. 34080–34087.
[140] T.C. Osborne, O. Obianyo, X. Zhang, X. Cheng, P.R. Thompson, Protein arginine [156] K. Arita, H. Hashimoto, T. Shimizu, K. Nakashima, M. Yamada, M. Sato, Structural
methyltransferase 1: positively charged residues in substrate peptides distal to basis for Ca(2+)-induced activation of human PAD4, Nat. Struct. Mol. Biol. 11
the site of methylation are important for substrate binding and catalysis, (2004) 777–783.
Biochemistry 46 (2007) 13370–13381. [157] P.L. Kearney, M. Bhatia, N.G. Jones, L. Yuan, M.C. Glascock, K.L. Catchings, M.
[141] S. Klein, J.A. Carroll, Y. Chen, M.F. Henry, P.A. Henry, I.E. Ortonowski, G. Pintucci, Yamada, P.R. Thompson, Kinetic characterization of protein arginine deiminase
R.C. Beavis, W.H. Burgess, D.B. Rifkin, Biochemical analysis of the arginine 4: a transcriptional corepressor implicated in the onset and progression of
methylation of high molecular weight fibroblast growth factor-2, J. Biol. Chem. rheumatoid arthritis, Biochemistry 44 (2005) 10570–10582.
275 (2000) 3150–3157. [158] Z. Dong, P. Saikumar, J.M. Weinberg, M.A. Venkatachalam, Calcium in cell injury
[142] M.A. Lischwe, R.G. Cook, Y.S. Ahn, L.C. Yeoman, H. Busch, Clustering of glycine and and death, Annu. Rev. Pathol. 1 (2006) 405–434.
NG,NG-dimethylarginine in nucleolar protein C23, Biochemistry 24 (1985) [159] P.R. Thompson, W. Fast, Histone citrullination by protein arginine deiminase: is
6025–6028. arginine methylation a green light or a roadblock? ACS Chem. Biol. 1 (2006)
[143] M.A. Lischwe, R.L. Ochs, R. Reddy, R.G. Cook, L.C. Yeoman, E.M. Tan, M. Reichlin, H. 433–441.
Busch, Purification and partial characterization of a nucleolar scleroderma [160] G.L. Cuthbert, S. Daujat, A.W. Snowden, H. Erdjument-Bromage, T. Hagiwara, M.
antigen (Mr = 34,000; pI, 8.5) rich in NG,NG-dimethylarginine, J. Biol. Chem. 260 Yamada, R. Schneider, P.D. Gregory, P. Tempst, A.J. Bannister, T. Kouzarides,
(1985) 14304–14310. Histone deimination antagonizes arginine methylation, Cell 118 (2004) 545–553.
[144] J.J. Smith, K.P. Rucknagel, A. Schierhorn, J. Tang, A. Nemeth, M. Linder, H.R. [161] Y. Wang, J. Wysocka, J. Sayegh, Y.H. Lee, J.R. Perlin, L. Leonelli, L.S. Sonbuchner,
Herschman, E. Wahle, Unusual sites of arginine methylation in Poly(A)-binding C.H. McDonald, R.G. Cook, Y. Dou, R.G. Roeder, S. Clarke, M.R. Stallcup, C.D.
protein II and in vitro methylation by protein arginine methyltransferases PRMT1 Allis, S.A. Coonrod, Human PAD4 regulates histone arginine methylation levels
and PRMT3, J. Biol. Chem. 274 (1999) 13229–13234. via demethylimination, Science 306 (2004) 279–283.
[145] N. Troffer-Charlier, V. Cura, P. Hassenboehler, D. Moras, J. Cavarelli, Functional [162] R. Raijmakers, A.J. Zendman, W.V. Egberts, E.R. Vossenaar, J. Raats, C. Soede-
insights from structures of coactivator-associated arginine methyltransferase 1 Huijbregts, F.P. Rutjes, P.A. van Veelen, J.W. Drijfhout, G.J. Pruijn, Methylation of
domains, EMBO J. 26 (2007) 4391–4401. arginine residues interferes with citrullination by peptidylarginine deiminases in
[146] V.H. Weiss, A.E. McBride, M.A. Soriano, D.J. Filman, P.A. Silver, J.M. Hogle, The vitro, J. Mol. Biol. 367 (2007) 1118–1129.
structure and oligomerization of the yeast arginine methyltransferase, Hmt1, [163] M.S. Cosgrove, C. Wolberger, How does the histone code work? Biochem. Cell.
Nat. Struct. Biol. 7 (2000) 1165–1171. Biol. 83 (2005) 468–476.
[147] T.B. Miranda, M. Miranda, A. Frankel, S. Clarke, PRMT7 is a member of the protein [164] T. Jenuwein, C.D. Allis, Translating the histone code, Science 293 (2001)
arginine methyltransferase family with a distinct substrate specificity, J. Biol. 1074–1080.
Chem. 279 (2004) 22902–22907. [165] S.L. Schreiber, B.E. Bernstein, Signaling network model of chromatin, Cell 111
[148] K. Higashimoto, P. Kuhn, D. Desai, X. Cheng, W. Xu, Phosphorylation-mediated (2002) 771–778.
inactivation of coactivator-associated arginine methyltransferase 1, Proc. Natl. [166] Y. Tang, M.A. Holbert, H. Wurtele, K. Meeth, W. Rocha, M. Gharib, E. Jiang, P.
Acad. Sci. U. S. A. 104 (2007) 12318–12323. Thibault, A. Verrault, P.A. Cole, R. Marmorstein, Fungal Rtt109 histone
[149] J. Tang, J.D. Gary, S. Clarke, H.R. Herschman, PRMT 3, a type I protein arginine acetyltransferase is an unexpected structural homolog of metazoan p300/CBP,
N-methyltransferase that differs from PRMT1 in its oligomerization, subcellular Nat. Struct. Mol. Biol. 15 (2008) 738–745.

Das könnte Ihnen auch gefallen