Tải bản đầy đủ (.pdf) (11 trang)

Tài liệu Báo cáo khoa học: Physiological relevance of the endogenous mono(ADP-ribosyl)ation of cellular proteins doc

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (319.61 KB, 11 trang )

MINIREVIEW
Physiological relevance of the endogenous
mono(ADP-ribosyl)ation of cellular proteins
Maria Di Girolamo, Nadia Dani, Annalisa Stilla and Daniela Corda
Department of Cell Biology and Oncology, Consorzio Mario Negri Sud, Santa Maria Imbaro, Chieti, Italy
Enzyme-modulated mono(ADP-ribosyl)ation was ori-
ginally identified as the mechanism of action of several
of the bacterial toxins [1]. The diphtheria, cholera, per-
tussis and clostridia toxins are mono(ADP-ribo-
syl)transferases (ARTs; EC 2.4.2.31), and they are
known to cause various pathologies after their translo-
cation into mammalian host cells. Once inside the cell,
they act by modifying specific host cell proteins, such
as elongation factor 2, the a-subunit of the hetero-
trimeric GTP-binding (G) proteins, the small GTPases
Rho and Rac, and monomeric actin ([2–4], and refer-
ences therein).
More recently, a series of enzymes that are related to
these toxins have been identified in cells, and their
potential physiological roles have been explored ([5,6]
and references therein; Table 1). The best known of
these ARTs are ectoenzymes that are either glycosyl-
phosphatidylinositol (GPI)-anchored or secretory. Both
the toxins and these toxin-related eukaryotic ARTs
function through the transfer of an ADP-ribose residue
from bNAD
+
to a specific amino acid of the acceptor
protein, with the creation of an N- or S-glycosidic link-
age and the release of nicotinamide. The free amino acid
arginine (the most frequently modified residue of the


protein substrates) and its analogue agmatine have been
widely used as substrates to characterize the enzymatic
activities of these mono(ADP-ribosyl)transferases [7].
The mono(ADP-ribosyl)ation reaction is distinct from
Keywords
mono(ADP-ribosyl)ation; ADP-
ribosyltransferase; ART; G-protein; defensin;
apoptosis; P
2
X
7
Correspondence
D. Corda or M. Di Girolamo, Consorzio
Mario Negri Sud, Department of Cell
Biology and Oncology, 66030 Santa Maria
Imbaro (Chieti), Italy
Fax: +39 0872 570 412
Tel: +39 0872 570 338
E-mail: ,

Website: />(Received 18 April 2005, accepted 18 July
2005)
doi:10.1111/j.1742-4658.2005.04876.x
The mono(ADP-ribosyl)ation reaction is a post-translational modification
that is catalysed by both bacterial toxins and eukaryotic enzymes, and that
results in the transfer of ADP-ribose from bNAD
+
to various acceptor
proteins. In mammals, both intracellular and extracellular reactions have
been described; the latter are due to glycosylphosphatidylinositol-anchored

or secreted enzymes that are able to modify their targets, which include
the purinergic receptor P
2
X
7
, the defensins and the integrins. Intracellular
mono(ADP-ribosyl)ation modifies proteins that have roles in cell signalling
and metabolism, such as the chaperone GRP78 ⁄ BiP, the b-subunit of
heterotrimeric G-proteins and glutamate dehydrogenase. The molecular
identification of the intracellular enzymes, however, is still missing. A better
molecular understanding of this reaction will help in the full definition of
its role in cell physiology and pathology.
Abbreviations
ART, mono(ADP-ribosyl)transferase; ARTT, ADP-ribosylating turn-turn; DRAG, dinitrogenase reductase activating glycohydrolase; DRAT,
dinitrogenase reductase ADP-ribosyltransferase; FGF-2, basic fibroblast growth factor; G-protein, GTP binding protein; GPI, glycosyl-
phosphatidylinositol; IDDM, insulin-dependent diabetes mellitus; MIBG, meta-iodobenzylguanidine; NADase, NAD glycohydrolase; PARP,
poly(ADP-ribose) polymerases; PDGF, platelet-derived growth factor.
FEBS Journal 272 (2005) 4565–4575 ª 2005 FEBS 4565
that catalysed by the poly(ADP-ribose) polymerases
(PARPs; EC 2.4.2.30), which instead transfer branched
polymers of ADP-ribose to their target proteins, via
an O-glycosidic bond (reviewed in [7a]).
In mammalian cells, the mono(ADP-ribosyl)ation
reaction is also regulated by enzymes that are able
to reverse these post-translational modifications: the
cytosolic ADP-ribosyl hydrolases and the cytosolic and
extracellular pyrophosphatases [8,9]. With the former,
the protein-ADP-ribose linkage is hydrolysed to release
the ADP-ribose moiety, while with the pyrophospha-
tases, it is the pyrophosphate linkage that is hydro-

lysed, to release AMP and thus to leave a
ribosylphosphate attached to the protein.
The mammalian ecto-ADP-ribosyl-
transferases (ARTs)
The mammalian ARTs are coded for by a family of
structurally and functionally related genes. To date,
five mammalian enzymes (ART1–5) have been cloned,
although only four of these are expressed in humans,
due to a defective art2 gene that has a stop signal in
the coding region. Conversely, there are six ARTs
expressed in mouse, due to the duplication of the art2
gene ([5,6,10], and references therein).
This enzyme family shares very limited amino acid
sequence identity, with 20–30% seen among the ART
paralogue members within any species; the exception
here is mouse ART2.1 and ART2.2, where their
sequence identity (85%) indicates the recent evolution-
ary duplication of the mouse art2 gene [11]. The
ART2 enzymes have also been cloned from rat, and in
this case the two isoforms are known as ART2.a and
ART2.b and they are coded for by two alleles of a
single-copy gene [12]. These allelic differences between
the rat art2a and art2b genes result in a sequence
variation of only 10 amino acids between the ART2.a
and ART2.b proteins, although this alters their
enzymatic properties: while both can catalyze the
hydrolysis of NAD to ADP-ribose and nicotinamide,
only ART2.b is capable of auto-ADP-ribosylation [13].
As the human and mouse genome sequences have been
completely determined, all of the recognizable mem-

bers of these toxin-related ARTs have now been identi-
fied for these two species, with the identity among
orthologues ranging from 75% to 85%. As an exam-
ple, the deduced amino-acid sequence of mouse ART1,
which was the first cloned and characterized mamma-
lian arginine-specific ART, is 77% and 73% identical
to human and rabbit ART1, respectively [11,14].
Despite the low similarity at the level of their
amino-acid sequences, there are common structural
features that characterize this family of mammalian
ARTs [15,16]. The catalytic domain of these enzymes
is completely coded for by a single exon in all of the
ARTs, and it contains a conserved glutamate residue
that has been demonstrated to be crucial for the cata-
lytic activity of the bacterial toxins and of ART1 and
ART2 (mouse and rat) by site-specific mutagenesis. In
ART1 from rabbit, even the conservative glutamate
240 to aspartate (E240D) substitution abolishes the
transfer of ADP-ribose to the arginine used as an
acceptor; the neighbouring E238 has also been shown
to be important for ADP-ribose transfer [17,18]. In
several of the ARTs, the replacement of this second
glutamate abolishes the ability of these transferases to
use arginine as acceptors, thus further supporting the
hypothesis that this region is involved in substrate
recognition [19]. According to the structural model
proposed by Rappuoli and colleagues [16], this cata-
lytic domain is composed of 70–100 amino acids and
consists of three regions. Region 1, which is near the
N-terminal portion of the protein and is characterized

either by a conserved histidine (as in diphtheria toxin,
ART3 and the PARPs) or by a conserved arginine
(as in pertussis toxin, cholera toxin, the heat-labile
enterotoxins and the other ARTs); region 2, which is
Table 1. Mammalian ARTs. See text for details and relevant references.
Enzymes Source Substrate Effect of the reaction
Ectoenzymes
ART1 Human, rat, mouse Integrin, defensin, FGF-2, PDGFBB ⁄ Arg Inhibits substrate activity
ART2 Rat (a,b), mouse (1,2) P2X7, LFA1 ⁄ Arg Role in T-cell poliferation, apoptosis
ART3 Human, rat, mouse Unknown Unknown
ART4 Human, rat, mouse Unknown Unknown
ART5 Human, rat Unknown ⁄ Arg Unknown
Endoenzymes
Sirtuin2 Human Albumin ⁄ acetyl-lysine Involved in histone deacetylation
Arginine-specific Hamster, human Gb ⁄ Arg129 Inhibits substrate activity
Cysteine-specific Human GDH ⁄ Cys119 Inhibits substrate activity
mono(ADP-ribosyl)ation reactions in mammals M. Di Girolamo et al.
4566 FEBS Journal 272 (2005) 4565–4575 ª 2005 FEBS
characterized either by hydrophobic amino acids that
are involved in NAD
+
binding or by the serine-x-ser-
ine motif (where x represents threonine, serine or alan-
ine); and region 3, which is highly acidic and is
characterized by the conserved glutamate residue. The
arginine-serine-glutamate-x-glutamate motif (R-S-EQE;
which spans regions 1–2–3, respectively) is present in
cholera toxin and in ART1, 2 and 5, and it is typical
of the arginine-specific ARTs. This motif is missing in
ART3 and ART4 [20,21]. Through a comparative ana-

lysis of crystallographic structures, Han and Tainer
[19] have more recently extended the significance of the
region 3 sequences by identifying an ADP-ribosylating
turn-turn (ARTT) motif that they have implicated in
substrate recognition. Consistent with the relevance of
the ARTT motif, it has been shown that the auto-
ADP-ribosylation of ART2.b is abolished by muta-
tions of its R204, which is part of the ARTT motif.
Similarly, if the Y204 of ART2.a is mutated to an
arginine (Y204R), it is possible to promote ART2.a
auto-ADP-ribosylation [22].
The other common structural features of the ARTs
include an a-helix-rich N-terminal region, which repre-
sents a signal sequence for extracellular proteins, and a
C-terminal region folded into b-sheets, which is char-
acteristic of GPI-anchored membrane proteins [23]. As
mammalian (human and mouse) ART5 does not con-
tain this hydrophobic C-terminal signal sequence, it is
a secreted protein [24]. Finally, there are four cysteines
involved in disulfide bridge formation that are con-
served among all of the ART isoforms. Thus, accord-
ing to the rat ART2.2 crystal structure, C21 (C53 in
mART1) forms a disulfide bond with another con-
served cysteine at 223 (C272 in mART1) at the C-ter-
minus of the molecule, which stabilizes the folding of
the a-helix-rich domain [11]. C121 and C173 form the
second disulfide bond, which is also located at the pro-
tein surface and is also important for protein stabiliza-
tion [25].
The possibility that the N- and C-terminal domains

of the ARTs are involved in the regulation of ART
activity has been recently investigated by measuring
both the transferase and NAD glycohydrolase
(NADase) activities of truncated mutants of ART1
[26]. In mouse ART1, the amino acids at 24–38 (an
ART1-specific extension) modulate both the trans-
ferase and NADase activities, and amino acids 39–45
(a common ART coil) are essential for both activities.
The removal of the C-terminal basic domain decreases
the transferase, but enhances the NADase activity. The
N- and C-terminal regions of ART1 are therefore
required for its transferase activity, while the enhanced
NADase activity of the shorter mutants indicates that
there are sequences outside of the catalytic site that
exert structural constraints, and that modulate the sub-
strate specificity and catalytic activity [26].
The ecto-ARTs: expression and
function
ART1 is predominantly expressed in skeletal muscle,
heart and lung, and in neutrophils and T-cell lym-
phoma cells [27,28]. Its substrates include integrin a7
in mouse skeletal muscle cells, where ADP-ribosylation
has been proposed to have a role in myogenesis, as an
increase in arginine-specific ADP-ribosylation has been
observed during their differentiation into myotubes
[29].
Of note, the human defensin HNP-1 is among the
most recently identified substrates of ART1 [30]. The
defensins are 2–6 kDa cationic peptides that are con-
sidered to be the major components of innate anti-

microbial immunity, and that are thought to act by
disrupting the microbial membrane [31]. They can
also be considered to be components of adaptive
immunity, because cytokine stimulation of human
natural killer cells and T- and B-lymphocytes leads to
the production of the defensins. Interestingly, a-defen-
sins 1–3 are secreted by CD8 T-cells from immuno-
logically stable HIV-1-infected individuals (long-term
nonprogressors) and they are able to suppress HIV-1
replication [32].
Thus ART1 has been shown to modify HNP-1 on
R14 in an in vitro assay [30]. This ADP-ribosylated
HNP-1 loses its antimicrobial and cytotoxic activity,
although it significantly increases the release of IL-8
from A549 cells, as compared to unmodified HNP-1.
Conversely, the two peptides (unmodified and ADP-
ribosylated) have similar chemotactic activities when
evaluated for their ability to recruit T-lymphocytes
[30]. These data are consistent with the concept that,
once modified, HNP-1 acquires specific biological
activities that can result in the recruitment of neutro-
phils (by the release of IL-8 from epithelial cells) and
in the modulation of its own antimicrobial and cyto-
toxic activities. These latter aspects are particularly
relevant, as this study also identified ADP-ribosylated
HNP-1 in the bronchoalveolar lavage fluid from smok-
ers (but not from nonsmokers); this would indicate
that ADP-ribosylated HNP-1 is produced during the
inflammatory response (and loses its antimicrobial
activity). The relevance of these data also resides in the

fact that this was the first demonstration of endo-
genous ADP-ribosylation in humans [30,33].
Additional substrates of ART1 have been identified
in various different cell lines overexpressing this ecto-
M. Di Girolamo et al. mono(ADP-ribosyl)ation reactions in mammals
FEBS Journal 272 (2005) 4565–4575 ª 2005 FEBS 4567
enzyme, and these include growth factors and mem-
brane receptors. ART1-transfected rat adenocarcinoma
(NMU) cells were used to demonstrate ADP-ribosyla-
tion of basic fibroblast growth factor (FGF-2), which
had been detected initially on the surface of adult
bovine aortic endothelial and human hepatoma cells
[34,35]. As FGF-2 has a high affinity for heparin, it is
localized and possibly sequestered by the heparin sul-
fates on the cell surface and in the extracellular matrix.
Heparin also inhibits the ADP-ribosylation reaction,
which would imply that the heparin binding of FGF-2
and its ADP-ribosylation are mutually exclusive. Fur-
thermore, the ADP-ribosylated site of FGF-2 is in its
receptor-binding domain, and so it is possible that
ADP-ribosylation modulates the binding of FGF-2 to
its receptor and to heparin, thus regulating its avail-
ability to the cell [34,35].
In ART1-transfected V79 Chinese hamster lung
fibroblasts, platelet-derived growth factor-BB (PDGF-
BB) is the best substrate for ART1, whereas its struc-
tural homologue PDGF-AA is not a substrate [36].
ADP-ribosylated PDGF-BB loses its ability to stimu-
late mitogenic and chemotactic responses in human
pulmonary smooth muscle cells, and it shows a

reduced capacity for binding to PDGF receptors in
competition-binding experiments, as compared to
unmodified PDGF-BB [36]. This indicates that PDGF-
BB-dependent signalling can be regulated by ART1
activity at the cell surface.
When the EL-4 mouse T-cell lymphoma cell line was
stably transfected with ART1, the T-cell receptor sig-
nalling was inhibited in the presence of NAD via the
ADP-ribosylation of integrin LFA-1 and other co-
receptor proteins [37,38]. These effects have been pro-
posed to result from a failure of the T-cell receptors
and coreceptors to associate into a functional receptor
cluster. Thus these T-cell responses would be modula-
ted by mono(ADP-ribosyl)ation of cell surface proteins
[37,38].
In general, the role of the ARTs in T-cell signalling is
not clear, although the expression of ART2 on T-cells
has been well characterized. Thus ART2 is known to be
expressed in resting T-cells and in natural killer cells,
and it appears to be specific to the immune system. This
presence of ART2 on the surface of immune cells would
thus suggest an immunomodulatory activity, and
indeed, a significant disposition to develop autoimmune
diabetes has been shown to depend on the absence of
ART2 expression on rat T-cells [39–42].
What can perhaps be defined as the most intriguing
function of ART2 was recently uncovered by
Koch-Nolte and coworkers, namely that ADP-ribosy-
lation activates the P
2

X
7
purinoceptor ([43], Fig. 1A).
P
2
X
7
is a member of the P
2
X family of ATP-gated ion
channels, and it is widely expressed on several types of
blood cells [44]. This specific purinoceptor has attrac-
ted interest because of its particular ability to induce
the formation of large membrane pores. Thus the acti-
vation of P
2
X
7
with millimolar concentrations of ATP
triggers calcium fluxes, phosphatidylserine exposure
and apoptosis [44]. These same effects are trig-
gered by NAD at micromolar concentrations via the
ADP-ribosylation of P
2
X
7
. However, these effects are
not seen in ART2-deficient T-cells, demonstrating that
the activation of P
2

X
7
by NAD is ART2-dependent
[43]. These data provide an explanation for the previ-
ous demonstrations that extracellular NAD induces
rapid apoptosis in naive T-cells by a mechanism invol-
ving ADP-ribosylation of cell surface molecules [45].
Altogether, these data show that not only are ART1
and ART2 expressed in cells of the immune system,
but also that these two arginine-specific ARTs have a
clear role in the regulation of the immune response.
However, it is somewhat disappointing that one of the
most interesting physiological roles has been defined
for an ART that is not expressed in human cells. Thus,
it is important to understand what the human counter-
part of mouse ART2 might be, and whether these
mouse T-cell effects can be extended to human cells.
The biological functions of ART3, ART4 and
ART5 remain poorly defined [20,24]. ART3 and
ART5 are strongly expressed in human testis, whereas
ART4 is preferentially expressed in human
lymphatic tissue. In human monocytes, the cell-surface
ADP-ribosylated proteins are modified on their cys-
teine residues, suggesting that ART3 and ART4 are
cysteine-specific ARTs [21]. This is consistent with the
observation that in in vitro assays neither of these two
ARTs displays arginine-specific enzymatic activity
when expressed in and purified from Sf9 insect cells. In
the same study, human ART5 was seen to be an argin-
ine-specific ART, unlike mouse ART5, which shows a

potent NADase activity [5].
A point that still needs to be clarified is the occur-
rence of the extracellular NAD
+
that is required to
sustain the ADP-ribosylation reaction. The steady-
state concentration of NAD
+
in the serum of healthy
individuals is around 0.1 lm, and it can be kept low
by the extracellular NAD-glycohydrolase CD38 (both
soluble and membrane-associated) [46]; thus, to be util-
ized by ecto-ARTs, extracellular NAD
+
should reach
the concentration of 1–10 lm that is required for
ADP-ribosylation of P
2
X
7
[43], or higher if the K
m
of
the ARTs (from in vitro assays) is considered [47]. The
probable mechanism is that NAD
+
is released from
cells, where its concentration is in the range of
mono(ADP-ribosyl)ation reactions in mammals M. Di Girolamo et al.
4568 FEBS Journal 272 (2005) 4565–4575 ª 2005 FEBS

0.5–1.0 mm, as a consequence either of cell lysis during
inflammatory immune reactions and apoptosis, or of
nonlytic release, for example through the connexin 43
channels [48].
Intracellular mono(ADP-ribosyl)tran-
ferases and endogenous substrates
Although the ARTs that are able to modify extracellu-
lar proteins are the only well characterized family,
mono(ADP-ribosyl)ation has also been demonstrated
for intracellular proteins involved in cell signalling and
metabolism (Table 1; [6] and references therein). The
enzymatic activities involved here have been shown to
be both cytosolic and membrane associated, although
there is very little further information available
concerning their identities. The first example of a
well defined intracellular ADP-ribosylation cycle was
reported in prokaryotes. An intracellular ART activ-
ity (dinitrogenase reductase ADP-ribosyltransferase;
DRAT) was characterized in the photosynthetic bac-
terium Rodospirillum rubrum [49], where it regulates
nitrogen fixation through mono(ADP-ribosyl)ation on
R101 of the dinitrogenase reductase [50]. This reaction
is reversible and the dinitrogenase reductase is fully
reactivated by an ADP-ribosylarginine-hydrolase
known as dinitrogenase reductase activating glyco-
hydrolase (DRAG). Surprisingly, there is no significant
amino acid sequence similarity between DRAT and
the bacterial toxins that have ADP-ribosyltransferase
activity; only a few key residues are conserved across
the two families [23].

The same scenario could occur for the two families
of mammalian ADP-ribosyltransferases: the ecto-ARTs
and the endo-ARTs. These endo-ARTs appear to be
part of a completely different family of proteins that
A
B
Fig. 1. (A) Schematic representation of the
mammalian mono(ADP-ribosyl)ation reac-
tions. The figure shows both the extracellular
mono(ADP-ribosyl)ation that is catalysed by
the ARTs and the intracellular reaction that is
catalysed by the yet undefined mono(ADP-
ribosyl)transferases. The upper section
(extracellular space, OUT) shows the ART2-
dependent ADP-ribosylation of the P
2
X
7
purinergic receptor. The ADP-ribosylated
receptor is activated and leads to T-cell apop-
tosis. The lower section (intracellular space,
IN) shows the ADP-ribosylation ⁄ deribosyla-
tion cycle of the heterotrimeric G-protein b
subunit that is catalysed by a membrane-
associated, intracellular ADP-ribosyltransf-
erase (iART) and by a cytosolic ADP-ribo-
sylhydrolase (ARH). The dashed arrow
indicates possible hormonal regulation of this
iART. The effectors that are uncoupled from
the bc dimer by ADP-ribosylation are indica-

ted by the red line, while the red arrow indi-
cates coupling (see text for details). (B)
Schematic representation of the product of a
mono(ADP-ribosyl)ation reaction. The N-gly-
cosidic linkage between the ADP-ribose
residue and Arg129 on the heterotrimeric
G-protein b-subunit is illustrated.
M. Di Girolamo et al. mono(ADP-ribosyl)ation reactions in mammals
FEBS Journal 272 (2005) 4565–4575 ª 2005 FEBS 4569
shows no structural relationship to the ecto-ARTs des-
cribed above. An example consistent with this is seen
in the sirtuin family. It has recently been shown that
yeast silent information regulator 2 protein (Sir2p, a
NAD
+
-dependent histone ⁄ protein deacetylase) has
ADP-ribosyltransferase activity, and while it deacety-
lates histones, it also catalyses the mono(ADP-ribo-
syl)ation of the removed acetyl group [51,52]. Clearly,
this is not a ‘classical’ reaction that involves the modi-
fication of a target protein, but it involves small mole-
cule substrates. It is therefore similar to that seen for a
bacterial ADP-ribosyltransferase that is able to ADP-
ribosylate and inactivate the antibiotic rifampicin [53],
and a yeast enzyme that is able to transfer ADP-ribose
from NAD
+
to a phosphate group in tRNA [54]. To
date, seven human homologues of Sir2p have been des-
cribed (sirtuins 1–7), and they are characterized by

ART activity [51,52], although they share no obvious
sequence homologies with the ARTs themselves. Obvi-
ously, the sirtuins could represent the prototypes of a
novel intracellular ART family. The alternative possi-
bility that the ectocellular ARTs can modify intracellu-
lar substrates can also be considered. In this situation,
either there needs to be a search for new isoforms that
do not contain the signal peptide, or it needs to be
shown that one or more of the ecto-ARTs can be shed
from the membrane and can translocate into the cyto-
plasm. This could be achieved in a way similar to that
of the bacterial toxins, which have their own specific
receptors on the plasma membrane [55]. We are now
actively working to identify and define potential new
intracellular ART isoforms.
The intracellular mono(ADP-ribosyl)ation reactions
have been associated with cell signalling and metabolism
in intact cells. They modify three substrate proteins: the
endoplasmic reticulum-resident chaperone GRP78 ⁄ BiP,
the b-subunit of heterotrimeric G-proteins, and the
mitochondrial glutamate dehydrogenase GDH.
The mono(ADP-ribosyl)ation of GRP78 ⁄ BiP leads
to its inactivation. The modified GRP78 ⁄ BiP has been
detected in response to conditions that deplete the
endoplasmic reticulum of processible proteins or that
result in nutritional stress (such as lowered tempera-
ture, amino acid and glucose starvation), and has been
related to the rate of protein synthesis and processing
in intact Swiss 3T3 and GH3 pituitary cells [56–58].
According to the model proposed by Laitusis and col-

leagues [56], in cells with high rates of protein synthe-
sis, unmodified GRP78 ⁄ BiP is complexed with protein
folding intermediates; a slowing of protein synthesis
results in accumulation of the free, active form of
GRP78 ⁄ BiP, which is subjected to subsequent inactiva-
tion by ADP-ribosylation. The ADP-ribosylated form
of the chaperone thus provides a buffering system that
allows the rates of protein processing to be balanced
with those of protein synthesis. It should be noted that
while this mono(ADP-ribosyl)ation occurs intracellu-
larly, from a topological point of view the catalytic
domain of the enzyme involved (that has not yet been
characterized) needs to be located in the lumen of
the endoplasmic reticulum to modify its substrate,
GRP78 ⁄ BiP. Thus, this intracellular reaction occurs
out of the cytosolic compartment.
Direct evidence of functional, intracellular mono-
(ADP-ribosyl)ation has been reported for the G-protein
b-subunit ([59], Fig. 1). This reaction modifies R129 of
the b-subunit (Fig. 1B) and is catalysed by a
plasma-membrane-associated, but not GPI-anchored,
intracellular ART that has not yet been molecularly
characterized. The mono(ADP-ribosyl)ated b-subunit
becomes the substrate of a cytosolic, ADP-ribosylhydro-
lase ([59], Fig. 1), which completes a cellular ADP-
ribosylation ⁄ de-ADP-ribosylation cycle that controls
the activation ⁄ inactivation of the bc-dimer. Import-
antly, b-subunit mono(ADP-ribosyl)ation has also been
detected in intact cells, under both resting [59] and sti-
mulated conditions, thus indicating the physiological

potential of this reaction. In intact cells under resting
conditions, approximately 0.2% of the total bc-hetero-
dimer is modified; this could correspond to a cellular
pool of free bc-heterodimer that remains inactive. This
hypothesis is supported by the demonstration that the b-
subunit is modified only as a free heterodimer, and that
mono(ADP-ribosyl)ation inactivates the b-subunit by
impairing its interactions with its effector enzymes. This
has been shown directly in the case of type 1 adenylyl
cyclase, phosphoinositide 3-kinase and phospholipase
C [59,60]. Thus, the ADP-ribosylation ⁄ deribosylation
cycle modulates the function of the b-subunit. It is of
particular interest here that the ADP-ribosylation of the
b-subunit has also been shown to be under hormonal
control: it can be increased upon activation of specific
G-protein-coupled receptors (e.g. thrombin, serotonin
and cholecystokinin receptors), indicating that the active
bc-heterodimer released from different classes of G-pro-
teins can be a substrate for the endogenous mono(ADP-
ribosyl)transferase [60]. Thus, while activation of these
receptors will lead to the activation and dissociation of
the G-protein a- and the bc-subunits, this can initiate a
parallel inactivation of bc-subunit function that would
potentially regulate the duration of bc and a signalling,
through the selective termination of the bc function.
An ADP-ribosylation ⁄ deribosylation cycle has also
been proposed to occur in mitochondria, and involves
the cysteine-specific ADP-ribosylation of mitochondrial
GDH in intact Hep-G2 cells [61]. The modified
mono(ADP-ribosyl)ation reactions in mammals M. Di Girolamo et al.

4570 FEBS Journal 272 (2005) 4565–4575 ª 2005 FEBS
cysteine has been recently identified as Cys119 [61a].
As for GRP78 ⁄ BiP and for the G-protein b-subunit,
however, the nature of this ADP-ribosyltransferase
activity remains uncharacterized. This cycle appears to
be completed by an ADP-ribosylcysteine hydrolase
that is also present in mitochondria.
Further substrates of mono(ADP-ribosyl)ation have
also been identified, including the membrane-fissioning
protein CtBP3 ⁄ BARS [62,63], and the cytoskeletal pro-
teins actin, tubulin and desmin [64–66]. However, to
date there has been no direct evidence for their in vivo
modification. Overall, a better understanding of the
various and diverse biological roles of ADP-ribosyla-
tion of cellular proteins and peptides will be essential
to fully define the role of this modification in normal
and disease states.
The ADP-ribosylation reaction as a potential
new drug target
A lack of ART2 expression has been correlated with
an enhanced sensitivity to autoimmune disease in
several animal models [40,67]. For example, in dia-
betes-prone BioBreeding (DP-BB) rats, a model for
autoimmune insulin-dependent diabetes mellitus
(IDDM), a defective expression of ART2 in their
T-cells is associated with an increased susceptibility to
the disease [39–42]. Conversely, the prevention of
IDDM has been described in the same DP-BB rats fol-
lowing a transfusion with ART2-positive T-cells. The
development of IDDM has also been observed in dia-

betes-resistant BioBreeding (DR-BB) rats when they
are treated with a monoclonal antibody against ART2
[42]. Thus, ART2 expression appears to confer protec-
tion to IDDM in this animal model of the disease [42].
In the ART2.2 natural knock-out NZW mouse, the
development of a lupus-like glomerulonephritis has
been shown, again supporting the hypothesis that
ART2-positive T-cells confer protection against auto-
immune disease [68]. Polymorphisms have been repor-
ted for ART2.1 in the C57B1 ⁄ 6 mouse, where a stop
codon at position 481 leads to an ART enzyme that
lacks the GPI-anchor site and that has a reduced
transferase activity [69]. However, it is important to
stress that in these mouse models, disease development
is under the control of several genetic factors, and thus
a reduction or absence of ART2 expression is neces-
sary but not sufficient for the onset of autoimmune
pathologies. In line with this, both the NZW mouse
and other mouse models that are natural knock-outs
for ART2.1 (e.g. C57B1 ⁄ 6, BXSB), and the experi-
mentally induced ART2 double knock-out do not
show any evident immunological defects [70].
All of the data discussed above are consistent with the
hypothesis that NAD-induced cell death via the acti-
vation of the P
2
X
7
receptor has a role in immune
responses. However, as ART2 knock-out mice show

normal numbers and a normal distribution of T-cells
[70], this NAD-dependent cell death cannot be crucial in
the generation and maintenance of the T-cell. Rather, it
is possible that ART2-induced T-cell death has a role
during mechanical tissue injury or microbial inflamma-
tory processes with severe cytolysis. Under these circum-
stances, the massive release of intracellular antigens is
combined with high local concentrations of inflamma-
tory cytokines, raising the danger of activation of auto-
reactive T-cells. Thus ART2-induced T-cell death could
provide a safeguard mechanism against the undesirable
activation of irrelevant and potentially autoreactive
T-cells during an inflammatory response [43]. When
extended to the identification of the counterpart in
human cells, these findings open the exciting prospect of
using NAD and its metabolites to modify the function
of the P
2
X
7
receptor and other purinoceptors [43].
Other mechanisms that can benefit from ADP-
ribosylation-related drugs have emerged from a number
of recent reports. ADP-ribosylation has been coupled
to intracellular events that are associated with smooth
muscle cell vasoreactivity, cytoskeletal integrity and free
radical damage [71,72]. Additionally, there is evidence
that ADP-ribosylation is required for smooth muscle
cell proliferation [71,72]. Recent data have provided a
direct link between mono(ADP-ribosyl)ation and

smooth muscle cell proliferation and migration: meta-
iodobenzylguanidine (MIBG), a selective inhibitor of
arginine-dependent mono(ADP-ribosyl)ation, blocks
the stimulation of DNA and RNA synthesis, prevents
smooth muscle cell migration, and suppresses the induc-
tion of c-fos and c-myc gene expression. MIBG pro-
motes the phosphorylation of the Rho effector PRK1 ⁄ 2,
suggesting that mono(ADP-ribosyl)ation participates in
a Rho-dependent signalling pathway that is required for
immediate early gene expression. Furthermore, expres-
sion of the c-fos gene is the earliest proliferative event
that has shown sensitivity to MIBG treatment, and it
represents a novel mechanism by which mono(ADP-
ribosyl)ation can influence cellular processes [71,72].
As the heterotrimeric G-proteins have key roles in
cell regulation and the bc complex is essential in a
wide range of G-protein functions, including apoptosis,
chemotaxis, secretion and cell proliferation and
differentiation, we believe that our finding of the
mono(ADP-ribosyl)ation of the endogenous b-subunit
identifies a potential target for drug development.
Indeed, recent data have shown that cellular invasion
induced by src, met and leptin can be abrogated by
M. Di Girolamo et al. mono(ADP-ribosyl)ation reactions in mammals
FEBS Journal 272 (2005) 4565–4575 ª 2005 FEBS 4571
constitutively activated forms of the Gao ⁄ i subunits,
and can be induced by the coexpression of Gb1c2 [73].
Moreover, depletion of free Gbc heterodimers by
the C-terminus of the b adrenergic receptor kinase
(ct-bARK) results in a remarkable decrease in cellular

adhesion and spreading on a collagen matrix [74].
Thus Gbc dimers can be seen to be positive effectors
of invasion pathways that are induced by oncogenes
and epigenetic factors.
In line with the proposal that the ADP-ribosylated
defensins represent tools for the treatment of pulmonary
inflammation and other lung diseases [30,33], other
ADP-ribosylated peptides that mimic the modified por-
tions of the various ADP-ribosylation substrates and
inhibitors of the ADP-ribosylation reaction itself show
potential for the treatment of various pathologies, inclu-
ding autoimmune syndromes and proliferative diseases.
Acknowledgements
We wish to thank Dr C.P. Berrie for editorial assist-
ance, Ms. E. Fontana for preparation of the Figures
and the Italian Association for Cancer Research
(AIRC, Milano, Italy), Telethon, Italy (project
n. GGP030295) and the MIUR for financial support.
N.D. is supported by a fellowship from AIRC.
References
1 Ueda K & Hayaishi O (1985) ADP-ribosylation. Annu
Rev Biochem 54, 73–100.
2 Krueger KM & Barbieri JT (1995) The family of bacter-
ial ADP-ribosylating exotoxins. Clin Microbiol Rev 8,
34–47.
3 Pallen MJ, Lam AC, Loman NJ & McBride A (2001)
An abundance of bacterial ADP-ribosyltransferases –
implications for the origin of exotoxins and their human
homologues. Trends Microbiol 9, 302–307; discussion
308.

4 Di Girolamo M & Corda D (2003) Mono-ADP-ribosy-
lation of heterotrimeric G proteins. In Handbook of Cell
Signalling (Bradshaw R & Dennis E, eds), Vol. 2,
pp. 613–618. Academic Press, San Diego.
5 Glowacki G, Braren R, Firner K, Nissen M, Kuhl M,
Reche P, Bazan F, Cetkovic-Cvrlje M, Leiter E, Haag F
& Koch-Nolte F (2002) The family of toxin-related
ecto-ADP-ribosyltransferases in humans and the mouse.
Protein Sci 11, 1657–1670.
6 Corda D & Di Girolamo M (2003) Functional aspects
of protein mono-ADP-ribosylation. EMBO J 22, 1953–
1958.
7 Okazaki IJ & Moss J (1999) Characterization of glyco-
sylphosphatidylinositiol-anchored, secreted, and intracel-
lular vertebrate mono-ADP-ribosyltransferases. Annu
Rev Nutr 19, 485–509.
7a Bu
¨
rkle A (2005) Poly(ADP-ribose). The most elaborate
metabolite of NAD
+
. FEBS J 272, 4576–4589.
8 Moss J, Stanley SJ, Nightingale MS, Murtagh JJ Jr,
Monaco L, Mishima K, Chen HC, Williamson KC &
Tsai SC (1992) Molecular and immunological character-
ization of ADP-ribosylarginine hydrolases. J Biol Chem
267, 10481–10488.
9 Zolkiewska A & Moss J (1995) Processing of ADP-ribo-
sylated integrin alpha 7 in skeletal muscle myotubes.
J Biol Chem 270, 9227–9233.

10 Haag F, Koch-Nolte F, Kuhl M, Lorenzen S & Thiele
HG (1994) Premature stop codons inactivate the RT6
genes of the human and chimpanzee species. J Mol Biol
243, 537–546.
11 Braren R, Glowacki G, Nissen M, Haag F & Koch-
Nolte F (1998) Molecular characterization and expres-
sion of the gene for mouse NAD+: arginine ecto-
mono(ADP-ribosyl) transferase, Art1. Biochem J 336,
561–568.
12 Bortell R, Kanaitsuka T, Stevens LA, Moss J, Mordes
JP, Rossini AA & Greiner DL (1999) The RT6 (Art2)
family of ADP-ribosyltransferases in rat and mouse.
Mol Cell Biochem 193, 61–68.
13 Haag F, Andresen V, Karsten S, Koch-Nolte F &
Thiele H (1995) Both allelic forms of the rat T cell dif-
ferentiation marker RT6 display nicotinamide adenine
dinucleotide (NAD)-glycohydrolase activity, yet only
RT6.2 is capable of automodification upon incubation
with NAD. Eur J Immunol 25, 2355–2361.
14 Okazaki IJ, Zolkiewska A, Nightingale MS & Moss J
(1994) Immunological and structural conservation of
mammalian skeletal muscle glycosylphosphatidylino-
sitol-linked ADP-ribosyltransferases. Biochemistry 33,
12828–12836.
15 Domenighini M, Magagnoli C, Pizza M & Rappuoli R
(1994) Common features of the NAD-binding and cata-
lytic site of ADP- ribosylating toxins. Mol Microbiol 14,
41–50.
16 Domenighini M & Rappuoli R (1996) Three conserved
consensus sequences identify the NAD-binding site of

ADP-ribosylating enzymes, expressed by eukaryotes,
bacteria and T-even bacteriophages. Mol Microbiol 21,
667–674.
17 Takada T, Iida K & Moss J (1995) Conservation of a
common motif in enzymes catalyzing ADP-ribose trans-
fer. Identification of domains in mammalian trans-
ferases. J Biol Chem 270, 541–544.
18 Hara N, Tsuchiya M & Shimoyama M (1996) Glutamic
acid 207 in rodent T-cell RT6 antigens is essential for
arginine-specific ADP-ribosylation. J Biol Chem 271,
29552–29555.
19 Han S & Tainer JA (2002) The ARTT motif and a uni-
fied structural understanding of substrate recognition in
mono(ADP-ribosyl)ation reactions in mammals M. Di Girolamo et al.
4572 FEBS Journal 272 (2005) 4565–4575 ª 2005 FEBS
ADP-ribosylating bacterial toxins and eukaryotic ADP-
ribosyltransferases. Int J Med Microbiol 291 , 523–529.
20 Koch-Nolte F, Haag F, Braren R, Kuhl M, Hoovers J,
Balasubramanian S, Bazan F & Thiele HG (1997) Two
novel human members of an emerging mammalian gene
family related to mono-ADP-ribosylating bacterial
toxins. Genomics 39, 370–376.
21 Grahnert A, Friedrich M, Pfister M, Haag F, Koch-
Nolte F & Hauschildt S (2002) Mono-ADP-ribosyl-
transferases in human monocytes: regulation by lipo-
polysaccharide. Biochem J 362, 717–723.
22 Stevens LA, Bourgeois C, Bortell R & Moss J (2003)
Regulatory role of arginine 204 in the catalytic activity
of rat alloantigens ART2a and ART2b. J Biol Chem
278, 19591–19596.

23 Bazan JF & Koch-Nolte F (1997) Sequence and struc-
tural links between distant ADP-ribosyltransferase
families. Adv Exp Med Biol 419 , 99–107.
24 Glowacki G, Braren R, Cetkovic-Cvrlje M, Leiter EH,
Haag F & Koch-Nolte F (2001) Structure, chromosomal
localization, and expression of the gene for mouse ecto-
mono (ADP-ribosyl) transferase ART5. Gene 275,
267–277.
25 Mueller-Dieckmann C, Ritter H, Haag F, Koch-Nolte
F & Schulz GE (2002) Structure of the ecto-ADP-ribo-
syl transferase ART2.2 from rat. J Mol Biol 322,
687–696.
26 Bourgeois C, Okazaki I, Cavanaugh E, Nightingale M
& Moss J (2003) Identification of regulatory domains in
ADP-ribosyltransferase-1 that determine transferase and
NAD glycohydrolase activities. J Biol Chem 278,
26351–26355.
27 Peterson JE, Larew JS & Graves DJ (1990) Purification
and partial characterization of arginine-specific ADP-
ribosyltransferase from skeletal muscle microsomal
membranes. J Biol Chem 265, 17062–17069.
28 Zolkiewska A, Nightingale MS & Moss J (1992) Mole-
cular characterization of NAD: arginine ADP-ribosyl-
transferase from rabbit skeletal muscle. Proc Natl Acad
Sci USA 89, 11352–11356.
29 Zolkiewska A & Moss J (1993) Integrin alpha 7 as sub-
strate for a glycosylphosphatidylinositol- anchored
ADP-ribosyltransferase on the surface of skeletal muscle
cell. J Biol Chem 268, 25273–25276.
30 Paone G, Wada A, Stevens LA, Matin A, Hirayama T,

Levine RL & Moss J (2002) ADP ribosylation of
human neutrophil peptide-1 regulates its biological
properties. Proc Natl Acad Sci USA 99, 8231–8235.
31 Yang D, Biragyn A, Kwak LW & Oppenheim JJ (2002)
Mammalian defensins in immunity: more than just
microbicidal. Trends Immunol 23, 291–296.
32 Zhang L, Yu W, He T, Yu J, Caffrey RE, Dalmasso
EA, Fu S, Pham T, Mei J, Ho JJ, Zhang W, Lopez P &
Ho DD (2002) Contribution of human alpha-defensin 1,
2, and 3 to the anti-HIV-1 activity of CD8 antiviral fac-
tor. Science 298, 995–1000.
33 Corda D & Di Girolamo M (2002) Mono-ADP-ribosy-
lation: a tool for modulating immune response and cell
signaling. Sci STKE 2002, PE53.
34 Jones EM & Baird A (1997) Cell-surface ADP-ribosyla-
tion of fibroblast growth factor-2 by an arginine-specific
ADP-ribosyltransferase. Biochem J 323, 173–177.
35 Boulle N, Jones EM, Auguste P & Baird A (1995) Ade-
nosine diphosphate ribosylation of fibroblast growth
factor-2. Mol Endocrinol 9, 767–775.
36 Saxty BA, Yadollahi-Farsani M, Upton PD, Johnstone
SR & MacDermot J (2001) Inactivation of platelet-
derived growth factor-BB following modification by
ADP-ribosyltransferase. Br J Pharmacol 133, 1219–
1226.
37 Wang J, Nemoto E & Dennert G (1996) Regulation of
CTL by ecto-nictinamide adenine dinucleotide (NAD)
involves ADP-ribosylation of a p56lck-associated pro-
tein. J Immunol 156, 2819–2827.
38 Liu ZX, Yu Y & Dennert G (1999) A cell surface ADP-

ribosyltransferase modulates T cell receptor association
and signaling. J Biol Chem 274, 17399–17401.
39 Burstein D, Mordes JP, Greiner DL, Stein D, Naka-
mura N, Handler ES & Rossini AA (1989) Prevention
of diabetes in BB ⁄ Wor rat by single transfusion of
spleen cells. Parameters that affect degree of protection.
Diabetes 38, 24–30.
40 Greiner DL, Mordes JP, Handler ES, Angelillo M,
Nakamura N & Rossini AA (1987) Depletion of
RT6.1+ T lymphocytes induces diabetes in resistant
biobreeding ⁄ Worcester (BB ⁄ W) rats. J Exp Med 166,
461–475.
41 Greiner DL, Handler ES, Nakano K, Mordes JP &
Rossini AA (1986) Absence of the RT-6 T cell subset in
diabetes-prone BB ⁄ W rats. J Immunol 136, 148–151.
42 Whalen BJ, Greiner DL, Mordes JP & Rossini AA
(1994) Adoptive transfer of autoimmune diabetes melli-
tus to athymic rats: synergy of CD4+ and CD8+
T cells and prevention by RT6+ T cells. J Autoimmun
7, 819–831.
43 Seman M, Adriouch S, Scheuplein F, Krebs C, Freese
D, Glowacki G, Deterre P, Haag F & Koch-Nolte F
(2003) NAD-induced T cell death: ADP-ribosylation of
cell surface proteins by ART2 activates the cytolytic
P2X7 purinoceptor. Immunity 19, 571–582.
44 Di Virgilio F, Chiozzi P, Ferrari D, Falzoni S, Sanz
JM, Morelli A, Torboli M, Bolognesi G & Baricordi
OR (2001) Nucleotide receptors: an emerging family of
regulatory molecules in blood cells. Blood 97, 587–600.
45 Scheuplein F, Adriouch S, Glowacki G, Haag F, Seman

M & Koch-Nolte F (2003) Triggering of T-cell apopto-
sis by toxin-related ecto-ADP-ribosyltransferase ART2.
Ann NY Acad Sci 1010, 296–299.
M. Di Girolamo et al. mono(ADP-ribosyl)ation reactions in mammals
FEBS Journal 272 (2005) 4565–4575 ª 2005 FEBS 4573
46 Berthelier V, Tixier JM, Muller-Steffner H, Schuber F
& Deterre P (1998) Human CD38 is an authentic
NAD(P)+ glycohydrolase. Biochem J 330, 1383–1390.
47 Okazaki IJ, Kim HJ, McElvaney NG, Lesma E & Moss
J (1996) Molecular characterization of a glycosylphos-
phatidylinositol-linked ADP-ribosyltransferase from
lymphocytes. Blood 88, 915–921.
48 Bruzzone S, Franco L, Guida L, Zocchi E, Contini P,
Bisso A, Usai C & De Flora A (2001) A self-restricted
CD38-connexin 43 cross-talk affects NAD+ and cyclic
ADP-ribose metabolism and regulates intracellular cal-
cium in 3T3 fibroblasts. J Biol Chem 276, 48300–48308.
49 Ludden PW (1994) Reversible ADP-ribosylation as a
mechanism of enzyme regulation in procaryotes. Mol
Cell Biochem 138, 123–129.
50 Lowery RG & Ludden PW (1988) Purification and
properties of dinitrogenase reductase ADP-ribosyltrans-
ferase from the photosynthetic bacterium Rhodospirillum
rubrum. J Biol Chem 263, 16714–16719.
51 Frye RA (1999) Characterization of five human cDNAs
with homology to the yeast SIR2 gene: Sir2-like pro-
teins (sirtuins) metabolize NAD and may have protein
ADP-ribosyltransferase activity. Biochem Biophys Res
Commun 260, 273–279.
52 Frye RA (2000) Phylogenetic classification of prokaryo-

tic and eukaryotic Sir2-like proteins. Biochem Biophys
Res Commun 273, 793–798.
53 Quan S, Venter H & Dabbs ER (1997) Ribosylative
inactivation of rifampin by Mycobacterium smegmatis
is a principal contributor to its low susceptibility to this
antibiotic. Antimicrob Agents Chemother 41, 2456–2460.
54 Spinelli SL, Kierzek R, Turner DH & Phizicky EM
(1999) Transient ADP-ribosylation of a 2¢-phosphate
implicated in its removal from ligated tRNA during
splicing in yeast. J Biol Chem 274, 2637–2644.
55 Lencer WI, Hirst TR & Holmes RK (1999) Membrane
traffic and the cellular uptake of cholera toxin. Biochim
Biophys Acta 1450, 177–190.
56 Laitusis AL, Brostrom MA & Brostrom CO (1999) The
dynamic role of GRP78 ⁄ BiP in the coordination of
mRNA translation with protein processing. J Biol Chem
274, 486–493.
57 Leno GH & Ledford BE (1989) ADP-ribosylation of
the 78-kDa glucose-regulated protein during nutritional
stress. Eur J Biochem 186, 205–211.
58 Staddon JM, Bouzyk MM & Rozengurt E (1992) Inter-
conversion of GRP78 ⁄ BiP. A novel event in the action
of Pasteurella multocida toxin, bombesin, and platelet-
derived growth factor. J Biol Chem 267, 25239–25245.
59 Lupi R, Corda D & Di Girolamo M (2000) Endogenous
ADP-ribosylation of the G protein beta subunit pre-
vents the inhibition of type 1 adenylyl cyclase. J Biol
Chem 275, 9418–9424.
60 Lupi R, Dani N, Dietrich A, Marchegiani A, Turacchio
S, Berrie CP, Moss J, Gierschik P, Corda D & Di Giro-

lamo M (2002) Endogenous mono-ADP-ribosylation of
the free Gbetagamma prevents stimulation of phospho-
inositide 3 kinase-gamma and phospholipase C- beta2
and Is activated by G-protein-coupled receptors.
Biochem J 367, 825–832.
61 Herrero-Yraola A, Bakhit SM, Franke P, Weise C,
Schweiger M, Jorcke D & Ziegler M (2001) Regulation
of glutamate dehydrogenase by reversible ADP-ribosyla-
tion in mitochondria. EMBO J 20 , 2404–2412.
61a Choi MM, Huh JW, Yang SJ, Cho EH, Choi SY &
Cho SW (2005) Identification of ADP-ribosylation site
in human glutamate dehydrogenase isozymes. FEBS
Lett 579, 4125–4130.
62 Di Girolamo M, Silletta MG, De Matteis MA, Braca
A, Colanzi A, Pawlak D, Rasenick MM, Luini A &
Corda D (1995) Evidence that the 50-kDa substrate of
brefeldin A-dependent ADP- ribosylation binds GTP
and is modulated by the G-protein beta gamma subunit
complex. Proc Natl Acad Sci USA 92, 7065–7069.
63 De Matteis MA, Di Girolamo M, Colanzi A, Pallas M,
Di Tullio G, McDonald LJ, Moss J, Santini G, Ban-
nykh S, Corda D & Luini A (1994) Stimulation of
endogenous ADP-ribosylation by brefeldin A. Proc Natl
Acad Sci USA 91, 1114–1118.
64 Zhou H, Huiatt TW, Robson RM, Sernett SW &
Graves DJ (1996) Characterization of ADP-ribosylation
sites on desmin and restoration of desmin intermediate
filament assembly by de-ADP-ribosylation. Arch Bio-
chem Biophys 334, 214–222.
65 Huang HY, Zhou H, Huiatt TW & Graves DJ (1996)

Target proteins for arginine-specific mono (ADP-ribo-
syl) transferase in membrane fractions from chick skele-
tal muscle cells. Exp Cell Res 226, 147–153.
66 Huang HY, Graves DJ, Robson RM & Huiatt TW
(1993) ADP-ribosylation of the intermediate filament
protein desmin and inhibition of desmin assembly
in vitro by muscle ADP-ribosyltransferase. Biochem
Biophys Res Commun 197, 570–577.
67 Ablamunits V, Bridgett M, Duffy T, Haag F, Nissen
M, Koch-Nolte F & Leiter H (2001) Changing patterns
of cell surface mono (ADP-ribosyl) transferase antigen
ART2.2 on resting versus cytopathically-activated
T cells in NOD ⁄ Lt mice. Diabetologia 44, 848–858.
68 Matthes M, Hollmann C, Bertuleit H, Kuhl M, Thiele
HG, Haag F & Koch-Nolte F (1997) ‘Natural’ RT6-1
and RT6-2 ‘knock-out’ mice. Adv Exp Med Biol 419,
271–274.
69 Kanaitsuka T, Bortell R, Stevens LA, Moss J, Sardinha
D, Rajan TV, Zipris D, Mordes JP, Greiner DL & Ros-
sini AA (1997) Expression in BALB ⁄ c and C57BL ⁄ 6
mice of Rt6–1 and Rt6–2 ADP-ribosyltransferases that
differ in enzymatic activity: C57BL ⁄ 6 Rt6–1 is a natural
transferase knockout. J Immunol 159, 2741–2749.
70 Ohlrogge W, Haag F, Lohler J, Seman M, Littman DR,
Killeen N & Koch-Nolte F (2002) Generation and
mono(ADP-ribosyl)ation reactions in mammals M. Di Girolamo et al.
4574 FEBS Journal 272 (2005) 4565–4575 ª 2005 FEBS
characterization of ecto-ADP-ribosyltransferase
ART2.1 ⁄ ART2.2-deficient mice. Mol Cell Biol 22,
7535–7542.

71 Yau L, Litchie B & Zahradka P (2004) MIBG, an inhi-
bitor of arginine-dependent mono(ADP-ribosyl)ation,
prevents differentiation of L6 skeletal myoblasts by inhi-
biting expression of myogenin and p21 (cip1). Exp Cell
Res 301, 320–330.
72 Yau L, Litchie B, Thomas S, Storie B, Yurkova N &
Zahradka P (2003) Endogenous mono-ADP-ribosylation
mediates smooth muscle cell proliferation and migration
via protein kinase N-dependent induction of c-fos
expression. Eur J Biochem 270, 101–110.
73 Faivre S, Regnauld K, Bruyneel E, Nguyen QD, Mareel
M, Emami S & Gespach C (2001) Suppression of cellu-
lar invasion by activated G-protein subunits Gao, Gai1,
Gai2, and Gai3 and sequestration of Gbc. Mol Pharma-
col 60, 363–372.
74 Bookout AL, Finney AE, Guo R, Peppel K, Koch WJ
& Daaka Y (2003) Targeting Gbetagamma signaling to
inhibit prostate tumor formation and growth. J Biol
Chem 278, 37569–37573.
M. Di Girolamo et al. mono(ADP-ribosyl)ation reactions in mammals
FEBS Journal 272 (2005) 4565–4575 ª 2005 FEBS 4575

×