ArticlePDF AvailableLiterature Review

Activation induced deaminase: How much and where?

Authors:

Figures

Major post-translational regulation steps affecting AID levels. (A) Schematic representation of steps participating in stability and subcellular localization regulation represented within a cycle. Only selected AID interacting factors are shown. AID is synthesized in the cytoplasm, where unfolded AID is met by the HSP40-HSC70 system, the specific action of the HSP40 DnaJa1 allows transferring AID into the HSP90 molecular chaperoning stabilization cycle. After some undefined maturation step, or conformational change, AID is passed onto eEF1 and/or other cytoplasmic retention factors before active nuclear import. A number of factors could be implicated, alternatively or jointly in AID nuclear import. Inside the nucleus AID is either exported by CRM1 or targeted to the Ig loci by interacting with a number of RNA processing factors, where it is phosphorylated by PKA. AID is (subsequently) degraded in the nucleus either through ubiquitin-or REG-dependent proteasomal degradation. (B) Simplified schematic representation of the same steps as in A, but in the form of a network in which most pools of AID are interconnected (see text). (C) Schematic representation of AID cytoplasmic degradation following inhibition of the HSP90 molecular chaperoning pathway. HSP90 inhibitors prevent the ATP hydrolysis cycle of the chaperone. FTI, farnesyltransferase inhibitors, prevent farnesylation of DnaJa1, which is required for binding to and stabilization of AID. Both inhibitors lead to polyubiquitination and proteasomal degradation of cytoplasmic AID.
… 
Content may be subject to copyright.
Seminars
in
Immunology
24 (2012) 246–
254
Contents
lists
available
at
SciVerse
ScienceDirect
Seminars
in
Immunology
j
ourna
l
ho
me
page:
www.elsevier.com/locate/ysmim
Review
Activation
induced
deaminase:
How
much
and
where?
Alexandre
Orthweina,b,
Javier
M.
Di
Noiaa,b,c,d,
aInstitut
de
Recherches
Cliniques
de
Montréal,
Montréal,
Québec,
H2W
1R7,
Canada
bDepartment
of
Microbiology
and
Immunology,
Université
de
Montréal,
Montréal,
Québec,
H3C
3J7,
Canada
cDepartment
of
Medicine,
Faculty
of
Medicine,
Université
de
Montréal,
Montréal,
Québec,
H3C
3J7,
Canada
dDepartment
of
Medicine,
McGill
University,
Montréal,
Québec,
Canada
a
r
t
i
c
l
e
i
n
f
o
Keywords:
Activation-induced
deaminase
(AID)
B-lymphocyte
Antibody
gene
diversification
Somatic
hypermutation
Class
switch
recombination
Humoral
immunity
Subcellular
localization
Protein
stability
a
b
s
t
r
a
c
t
Activation
induced
deaminase
(AID)
plays
a
central
role
in
adaptive
immunity
by
initiating
the
processes
of
somatic
hypermutation
(SHM)
and
class
switch
recombination
(CSR).
On
the
other
hand,
AID
also
pre-
disposes
to
lymphoma
and
plays
a
role
in
some
autoimmune
diseases,
for
which
reasons
AID
expression
and
activity
are
regulated
at
various
levels.
Post-translational
mechanisms
regulating
the
amount
and
subcellular
localization
of
AID
are
prominent
in
balancing
AID
physiological
and
pathological
functions
in
B
cells.
Mechanisms
regulating
AID
protein
levels
include
stabilizing
chaperones
in
the
cytoplasm
and
proteins
efficiently
targeting
AID
to
the
proteasome
within
the
nucleus.
Nuclear
export
and
cytoplasmic
retention
contribute
to
limit
the
amount
of
AID
accessing
the
genome.
Additionally,
a
number
of
factors
have
been
implicated
in
AID
active
nuclear
import.
We
review
these
intertwined
mechanisms
proposing
two
scenarios
in
which
they
could
interact
as
a
network
or
as
a
cycle
for
defining
the
optimal
amount
of
AID
protein.
We
also
comparatively
review
the
expression
levels
of
AID
necessary
for
its
function
during
the
immune
response,
present
in
different
cancers
as
well
as
in
those
tissues
in
which
AID
has
been
implicated
in
epigenetic
remodeling
of
the
genome
by
demethylating
DNA.
© 2012 Elsevier Ltd. All rights reserved.
1.
Antibody
diversification
in
germinal
center
B
cells
V(D)J
recombination
assembles
the
primary
repertoire
of
anti-
body
genes
during
B
cell
development.
B
cells
that
have
successfully
rearranged
their
immunoglobulin
(Ig)
genes
express
membrane
IgM
and/or
IgD
and
migrate
to
the
periphery
where
they
will
be
exposed
to
foreign
antigens.
The
first
cognate
antibody–antigen
recognition
of
naïve
B
cells
is
usually
not
of
high
affinity.
This
endows
the
system
with
enough
flexibility
to
interact
with
almost
any
possible
antigen;
however,
high
affinity
antibody–antigen
interactions
are
critical
for
neutralizing
or
disposing
of
antigens.
Thus,
there
are
mechanisms
to
further
change
the
antibodies.
B
cells
activated
by
cognate
antigen
initiate
the
germinal
center
reaction
[1].
Germinal
center
B
cells
divide
rapidly
while
diversifying
the
genes
encoding
for
the
heavy
and
light
antibody
chains.
Diversifi-
cation
occurs
through
the
introduction
of
point
mutations
in
the
variable
regions
of
the
IgH
and
IgL
by
the
mechanism
of
somatic
hypermutation
(SHM).
Antibody
variants
produced
by
SHM
are
selected
for
improved
antigen
recognition
by
antigen
presenting
Abbreviations:
AID,
activation
induced
deaminase;
ALL,
acute
lymphoblastic
leukemia;
CML,
chronic
myeloid
leukemia;
CSR,
class
switch
recombination;
DLBCL,
diffuse
large
B
cell
lymphoma;
Ph+,
Philadelphia
chromosome-positive;
NES,
nuclear
export
signal;
NLS,
nuclear
localization
signal;
SHM,
somatic
hypermutation.
Corresponding
author
at:
IRCM,
110
Avenue
des
Pins
Ouest,
Montréal,
Québec,
H2W
1R7,
Canada.
Tel.:
+1
514
987
5642;
fax:
+1
514
987
5528.
E-mail
address:
javier.di.noia@ircm.qc.ca
(J.M.
Di
Noia).
cells
and
T
cells,
which
leads
to
affinity
maturation
of
the
antibody
response.
At
the
same
time,
the
constant
exons
of
the
IgH
encod-
ing
for
IgM
and
IgD
are
exchanged
by
exons
encoding
IgG,
IgE,
or
IgA
isotypes
through
the
mechanism
of
class
switch
recombina-
tion
(CSR);
thereby
leading
to
the
production
of
antibodies
with
conserved
specificity
but
different
biological
properties.
Both,
SHM
and
CSR
are
mutagenic
mechanisms
that
despite
having
different
end
results
(i.e.
single
point
mutations
versus
a
chromosomal
dele-
tion
of
several
tens
of
kbp),
share
several
molecular
steps.
The
most
important
step
common
to
both
SHM
and
CSR
is
their
initiation
by
the
enzyme
Activation
induced
deaminase
(AID)
[2,3].
AID
is
part
of
the
AID/APOBEC
family
of
cytosine
deaminase-
related
enzymes,
most
of
which
have
the
unique
capacity
of
deaminating
deoxycytidine
in
single
stranded
DNA
thus
converting
it
into
deoxyuridine
[4].
This
is
already
a
mutagenic
lesion
caus-
ing
a
C:G
to
T:A
base
change
after
replication.
Processing
of
the
uracil
by
base
excision
and
mismatch
repair
enzymes
leads
to
the
broader
spectrum
of
point
mutations
characterizing
SHM,
and
to
DNA
double
strand
breaks,
which
are
necessary
intermediates
in
CSR
(reviewed
in
[5–7]).
2.
AID
levels
and
disease
2.1.
AID
deficiency
and
haploinsufficiency
As
expected
from
its
central
role
in
antibody
diversification,
loss
of
function
mutations
of
AID
cause
an
immunodeficiency
syndrome
1044-5323/$
see
front
matter ©
2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.smim.2012.05.001
A.
Orthwein,
J.M.
Di
Noia
/
Seminars
in
Immunology
24 (2012) 246–
254 247
characterized
by
the
absence
of
switched
isotypes,
recurrent
infec-
tions,
and
lymphoid
tissue
hyperplasia
[3],
which
is
recapitulated
in
Aicda/
mice
[2].
Aicda+/
mice
show
an
intermediate
degree
of
lymphoid
hyperplasia,
between
that
of
wt
and
Aicda/
mice
[8,9].
AID
haploinsufficient
mice
have
40%
of
AID
mRNA
and
pro-
tein
levels
compared
to
wt
[9,10]
and
several
groups
have
found
reduced
CSR
and
SHM
in
Aicda+/
mice
[8–11].
The
decrease
in
CSR
is
roughly
proportional
to
the
decrease
in
AID
mRNA
and
protein
levels
although
there
are
some
discrepancies
among
the
studies,
which
could
originate
from
using
different
mouse
strains
and/or
different
in
vitro
experimental
conditions
[12,13].
There
is
consis-
tency
between
studies
in
that
SHM
is
reduced
to
30%
of
the
wt
in
Peyer
patches
of
Aicda+/
[8,9],
with
some
more
variability
when
measured
elsewhere,
ranging
from
70%
of
wt
in
lymph
nodes
JH4
[8]
to
only
15%
of
wt
in
S!
region
of
B
cells
activated
ex
vivo
[10].
These
observations
suggest
that
AID
is
limiting
for
antibody
diversifica-
tion
although
it
is
not
the
only
factor
determining
the
efficiency
of
CSR
and
SHM.
In
vivo,
the
effect
is
highly
compensated
for
by
selection
[9,10]
and
it
is
unclear
how
compromised
the
immune
system
of
an
individual
with
reduced
AID
levels
would
be.
Judging
from
the
lack
of
clinical
symptoms
in
humans
carrying
only
one
functional
AID
allele
[3,14],
this
is
probably
not
a
major
issue
for
human
health.
However,
it
would
have
in
all
likelihood
been
disad-
vantageous
during
evolution.
The
previous
considerations
suggest
a
mechanism
by
which
the
minimal
levels
of
AID
were
selected.
The
pathological
side
effects
of
AID
probably
set
the
upper
limit
of
AID
expression.
2.2.
AID
levels
in
antibody
diversification
and
cancer
Consistent
with
the
view
that
AID
levels
are
limiting
for
anti-
body
diversification,
higher
levels
of
AID
protein
generally
translate
into
more
CSR
and
SHM,
and
are
accompanied
by
an
increased
inci-
dence
of
potentially
transforming
genomic
lesions.
Mouse
models
with
modified
AID
expression
levels
nicely
provided
evidence
for
this.
First,
several
transgenic
mice
overexpressing
AID
have
been
made,
all
of
which
show
increased
CSR
and
SHM
[15–18].
There
are
differences
in
the
magnitude
of
the
effect,
which
might
be
related
to
the
transgene
design.
The
two
transgenic
lines
in
which
AID
was
expressed
from
a
ubiquitous
promoter
showed
a
signifi-
cant
but
modest
increase
[15,16,18].
In
these
mice,
B
cells
express
high
levels
of
AID
throughout
their
development
and
some
counter-
selection
is
possible,
which
could
dampen
the
extent
of
the
effect.
In
contrast,
transgenic
AID
under
the
control
of
the
Ig"
enhancer
showed
maximal
expression
in
germinal
center
B
cells
and
in
this
case
there
was
a
very
high
increase
in
CSR
and
SHM
levels
[17].
Secondly,
boosting
AID
expression
in
B
cells
by
removing
the
neg-
ative
post-transcriptional
regulation
of
AID
by
miR-155
also
leads
to
higher
levels
of
CSR
[19,20].
Yet,
none
of
these
mice
showed
increased
levels
of
switched
serum
Ig,
but
this
is
probably
regulated
at
another
level
(selection,
homeostatic
proliferation),
and
does
not
necessarily
reflect
the
intrinsic
CSR
capacity
of
the
B
cells.
This
lack
of
correlation
between
in
vitro
CSR
and
serum
Ig
levels,
has
also
been
observed
in
MSH2-deficient
and
UNG-deficient
mice
[21–24].
Interestingly,
mice
in
which
higher
levels
of
AID
were
achieved
by
removing
the
miR-155
binding
sites
from
its
mRNA
showed
reduced
affinity
maturation,
but
no
differences
in
the
quantity
or
quality
of
SHM
at
the
Ig
variable
regions
[19].
This
would
suggest
that
expression
of
AID
above
physiological
levels
could
compro-
mise
B
cell
viability,
in
line
with
the
evidence
that
AID
limits
the
size
of
the
germinal
center
by
causing
B
cell
apoptosis
[25].
It
might
be
that
the
upper
limit
for
AID
physiological
expression
levels
could
be
influenced
by
the
increased
apoptosis
that
occurs
with
elevated
AID.
AID
also
contributes
to
the
development
of
cancer,
but
whether
AID
levels
correlate
proportionally
with
the
risk
of
developing
cancer
is
unclear.
AID
oncogenicity
is
most
likely
a
consequence
of
its
capacity
to
mutate
and
produce
DNA
breaks,
thus
initiating
chromosomal
translocations
affecting
a
number
of
loci
in
normal
B
cells
[26–30].
Increased
levels
of
AID
in
B
cells
correlates
with
more
IgH-cMyc
translocations
[17,20]
and
increased
mutations
in
some
non-Ig
targets
in
vivo
and
in
vitro
[17,19,31].
However,
this
does
not
always
have
oncogenic
consequences
[16,20]
(see
below).
Mice
overexpressing
AID
have
conclusively
demonstrated
the
oncogenic
capacity
of
AID.
Ubiquitous
transgenic
overexpression
of
AID
leads
to
T
cell
lymphomas
as
well
as
lung
adenomas
and
adeno-
carcinomas,
but
not
B
cell
malignancies
[18].
B
cells,
in
comparison
to
other
cell
types
that
do
not
express
AID,
may
have
evolved
pro-
tective
mechanisms
against
transformation,
thus
explaining
this
observation.
Indeed,
the
apoptotic
control
that
normally
eliminates
cells
with
AID-induced
translocations
[26],
prevented
B
cell
lym-
phomas
in
an
AID-Ig"transgenic
model
[17].
In
the
absence
of
p53,
AID-Ig"developed
predominantly
mature
B
cell
lymphomas,
even
outpacing
the
T
cell
lymphomas
that
usually
kill
p53-deficient
mice
[17].
Curiously,
germinal
center
B
cells
have
reduced
levels
of
p53
to
allow
for
the
necessary
DNA
damage
that
accompanies
antibody
gene
diversification
[32],
which
suggests
the
presence
of
additional
mechanisms
to
prevent
AID-initiated
lymphomagenesis
(for
DNA
repair
regulation
see
Saribasak
and
Gearhart,
this
issue).
For
instance,
off-target
AID
mutations
are
more
frequent
in
the
absence
of
a
number
of
DNA
repair
pathways,
even
with
normal
endogenous
AID
levels
[31,33,34].
Endogenous
levels
of
AID
do
initiate
B
cell
transformation,
albeit
infrequently.
The
etiological
role
of
AID
in
B
cell
lymphomas
orig-
inating
from
germinal
center
B
cells
was
demonstrated
using
the
I!HABCL6
transgenic
oncogene
model,
which
deregulates
BCL6
in
B
cells
and
results
in
frequent
mature
B
cell
lymphomas.
When
crossed
with
Aicda/
mice,
I!HABCL6
transgenic
mice
did
not
develop
mature
B
cell
lymphomas
[35].
I!HABCL6
lymphomas
are
akin
to
human
DLBCL
(diffuse
large
B
cell
lymphoma),
which
displays
a
high
prevalence
of
AID
expression
[36–40],
strongly
sug-
gesting
an
etiological
role
for
endogenous
AID
in
this
lymphoma.
In
addition,
endogenous
AID
levels
contribute
to
chromosomal
translocations
of
multiple
partners
with
the
Ig
locus,
such
as
the
IgH-Ig#
[41]
or
IgH-Pax5
[29,30],
and
even
between
two
non-Ig
genes
[17,29,30].
Indeed,
most
of
the
translocations
occurring
in
a
p53-deficient
background
occur
between
c-Myc
and
miR-142
[17],
a
molecular
hallmark
of
acute
prolymphocytic
leukemia
[42].
Although
a
few
DLBCL
samples
show
higher
than
normal
AID
levels,
in
general
DLBCL
and
follicular
lymphoma
samples
show
similar
or
lower
AID
levels
than
germinal
center
B
cells
[36–38,40].
Although
lymphoma
biopsies
contain
normal
and
non-B
cells,
where
only
AID
mRNA
has
been
measured
in
many
cases,
it
does
not
seem
that
AID
overexpression
is
necessary
for
lymphomagenesis.
The
latter
was
demonstrated
in
the
Balb/Bcl-xL
mouse
model
of
plasmacytoma.
In
this
model,
AID
is
required
for
most
pristane-induced
plasmacy-
tomas,
where
it
underpins
the
oncogenic
IgH-cMyc
translocation
[34,39].
AID
haploinsufficient
mice
show
reduced
incidence
of
lym-
phoma
compared
to
Aicda+/+,
but
still
significantly
higher
than
the
Aicda/
[10].
Thus,
significantly
lower
than
normal
levels
of
AID
are
sufficient
to
cause
plasmacytomas
when
they
are
combined
with
another
predisposing
condition.
The
IgH-cMyc
translocation,
which
is
caused
by
AID
[26,28,43,44],
is
a
hallmark
of
Burkitt’s
lym-
phoma
in
humans.
AID
is
expressed
at
near
normal
levels
in
this
type
of
lymphoma
[36].
Near
to
normal
AID
levels
are
also
expressed
in
Ph+
ALL
(Philadelphia
chromosome-positive
B
cell
acute
lym-
phoblastic
leukemia)
and
CML
(chronic
myelogenous
leukemia)
[45–47].
In
ALL
and
CML,
AID
can
be
a
disease
progression
factor
by
accelerating
leukemia
clonal
evolution
[46]
and/or
by
mutat-
ing
the
BCR-ABL1
oncogene,
thus
underpinning
resistance
to
the
therapeutic
drug
imatinib
[47].
Chronic
lymphocytic
leukemia
is
another
B
cell
malignancy
expressing
variable
levels
of
AID
due
248 A.
Orthwein,
J.M.
Di
Noia
/
Seminars
in
Immunology
24 (2012) 246–
254
to
population
heterogeneity,
where
only
a
defined
subpopulation
of
cells
expresses
most
of
AID
mRNA
[48,49].
The
presence
of
this
subpopulation
correlates
with
a
worst
prognosis,
but
AID
mRNA
in
these
cells
is
not
higher
than
in
normal
B
cells
[48,49].
Finally,
AID
is
also
expressed
in
a
number
of
tumor
cell
lines
and
primary
tumors
of
non-lymphoid
origin
including
breast
[50],
prostate
[51],
stomach
[52],
liver
[53,54]
and
lung
[55],
where
AID
is
normally
not
expressed
or
expressed
at
much
lower
levels
than
in
B
cells
[56].
2.3.
AID
and
autoimmunity
AID
levels
also
influence
antibody-mediated
autoimmune
dis-
eases.
AID
is
important
in
determining
the
severity
of
the
pathology,
as
demonstrated
in
mouse
models
of
lupus
and
arthritis
in
which
AID
deficiency,
or
even
haploinsufficiency,
results
in
a
less
severe
disease
[11,57–59].
In
keeping
with
this,
increased
levels
of
AID
cor-
relate
with
higher
levels
of
autoantibodies
in
the
MRL/faslpr/lpr and
BXD2
mouse
models
of
lupus
and
rheumatoid
arthritis,
respectively
[57,60–62].
Interestingly,
AID-deficient
mice
also
have
autoim-
mune
disorders
albeit
of
a
different
nature
[63].
This
fits
well
with
the
predisposition
to
autoimmune
disorders
noted
in
AID-deficient
human
patients
[64].
The
recent
finding
that
low
AID
expression
levels
during
B
cell
development
play
a
role
in
establishing
B
cell
tolerance
could
explain
these
findings
[65,66].
Although,
like
in
can-
cer,
there
are
predisposing
conditions
beyond
AID
for
developing
autoimmunity,
here
again
we
find
that
AID
levels
are
important
in
balancing
an
efficient
immune
response
with
disease.
This
equilib-
rium
may
be
altered
by
factors
inducing
AID,
such
as
estrogens
[67],
and
could
in
part
explain
the
higher
susceptibility
to
autoimmunity
in
women.
3.
Regulating
AID
The
data
summarized
above
illustrates
how
the
optimal
expres-
sion
levels
of
AID
in
germinal
center
B
cells
was
selected
during
vertebrate
evolution
as
a
compromise
between
being
able
to
mount
an
effective
adaptive
immune
response
and
delaying
the
onset
of
cancer
or
autoimmune
diseases.
A
strikingly
complex
network
of
regulatory
mechanisms
exists
to
ensure
that
the
optimal
amount
of
biologically
active
AID
protein
gets
to
the
Ig
locus.
Any
alteration
in
the
efficacy
of
these
mechanisms
could
predispose
to
immunodefi-
ciency,
autoimmunity
or
B
cell
lymphomas.
Increased
longevity
and
reduced
selective
pressure
in
humans
have
turned
evolutionary
irrelevant
side
effects
of
AID
into
medical
problems.
Understanding
AID
regulation
and
being
able
to
manipulate
its
levels
could
pro-
vide
ways
of
targeting
AID
to
prevent
or
ameliorate
some
of
these
pathologies.
3.1.
Aicda
expression
in
and
outside
B
cells
3.1.1.
Aicda
in
germinal
center
B
cells
and
cancer
Aicda
transcription
is
induced
by
cytokines
and
cell–cell
interac-
tions
in
the
context
of
antigen-triggered
B
cell
activation
during
an
immune
response
(reviewed
in
[68]
and
see
Vuong
and
Chaudhuri
this
issue).
Thus,
Aicda
is
highly
transcribed
in
germinal
center
cells
but
stringently
repressed
in
plasma
and
memory
B
cells
[69,70].
Measurements
of
AID
mRNA
in
the
B
cell
lineage
agree
well
with
western
blot
and
immunohistochemistry
data,
as
well
as
with
the
analysis
of
an
AID-GFP
reporter
mouse
[40,69,71,72].
A
combi-
nation
of
promoter,
enhancers
and
silencers
located
within
four
evolutionary-conserved
regulatory
regions
determines
AID
expres-
sion
in
germinal
center
B
cells
of
the
spleen
and
mucosal-associated
lymphoid
tissues
(Peyer’s
patches,
tonsils,
lymph
nodes)
[68,73].
However,
the
original
belief
of
AID
being
exclusively
expressed
in
germinal
center
B
cells
has
been
revised,
as
AID
expression
is
also
found
in
other
tissues,
not
only
in
several
pathological
situations
but
also
in
normal
tissues.
It
is
conceivable
that
the
expression
of
AID
in
B
lymphoid
malig-
nancies
can
be
induced
by
similar
stimuli
than
in
mature
B
cells,
or
remains
as
a
relict
of
the
physiological
counterpart
of
the
neo-
plasia.
In
other
cases,
such
as
leukemia
originating
from
progenitor
B
cells
(like
in
Ph+
ALL)
or
myeloid
cells
(like
CML),
where
AID
has
been
convincingly
detected
[45,47],
it
is
less
clear
but
the
BCR-ABL1
oncogene
itself
could
be
involved
in
inducing
Aicda
[45].
It
is
also
unclear
how
AID
expression
is
induced
in
malignant
epithelial
cells
from
solid
tumors.
Ectopic
AID
expression
in
these
tissues
is
likely
to
be
caused
by
the
transcription
factor
Nf-"B,
a
critical
transcrip-
tion
factor
in
orchestrating
inflammatory
and
innate
immunity
responses.
In
fact,
Nf-"B
participates
in
Aicda
induction
in
nor-
mal
B
cells
[44,74]
and
was
shown
to
mediate
AID
expression
in
hepatocytes
[52].
Furthemore,
epithelial
cancers
with
AID
expres-
sion
are
often
associated
to
chronic
inflammation,
sometimes
with
an
underlying
infection
such
as
Hepatitis
C
virus
[52,75],
Heli-
cobacter
pylori
[76,77]
or
human
immunodeficiency
virus
[78,79].
Also,
other
infectious
agents
like
Epstein-Barr
virus
[78],
Abelson
murine
leukemia
virus
[80]
and
Moloney
murine
leukemia
virus
[81],
or
stimulation
of
innate
immune
signaling
through
toll-like
receptors
[44]
can
induce
AID,
probably
through
Nf-"B
[81].
In
summary,
it
is
difficult
to
generalize
the
reasons
behind
AID
expres-
sion
in
cancer
cells,
let
alone
its
relevance
for
the
pathology.
It
is
likely
that
the
clinical
history
and
the
selection
acting
in
each
case
influence
the
variations
in
AID
expression
levels
we
see
between
patients.
3.1.2.
Aicda
transcription
in
other
normal
tissues
AID
is
expressed
outside
of
the
B
cell
compartment
under
nor-
mal
conditions.
The
most
notable
example
is
the
ovary,
in
which
basal
AID
mRNA
levels
are
50–70%
of
those
found
in
the
spleen
[56,82].
Furthermore,
Aicda
is
induced
by
estrogen
which
increase
AID
mRNA
a
few
fold
in
spleen
but
by
>20-fold
in
the
ovary
[67].
Breast
tissue
also
expresses
AID
when
stimulated
with
estrogens
[67],
and
several
human
breast
cancer
cells
express
AID
mRNA
[50],
but
we
do
not
know
the
basal
levels
in
normal
breast
tis-
sue.
Other
examples
of
AID
expression
in
normal
non-lymphoid
cells
include
prostate,
heart
and
lung,
although
the
estimates
of
AID
levels
in
these
cases
are
less
precise
[51,56,82].
The
presence
of
AID
would
suggest
a
physiological
role
for
it
in
these
tissues,
espe-
cially
in
the
ovary
where
levels
are
quite
high,
but
it
is
unclear
what
it
is.
Evidence
is
accumulating
to
suggest
that
AID
could
be
part
of
a
mechanism
to
demethylate
5meC
at
CpG
sites
in
the
genome,
thereby
influencing
gene
expression
[56,83,84].
A
role
in
epigenetic
reprograming
during
development
has
been
proposed,
which
could
explain
the
presence
of
AID
in
cell
types
such
as
oocytes,
spermato-
cytes,
primordial
germ
cells
or
embryonic
stem
cells
[56,67,82–86].
As
mentioned,
oocytes
express
AID
mRNA
to
comparable
levels
of
B
cells,
although
protein
expression
has
not
been
tested
[56,82].
AID
mRNA
is
very
low
in
testis
[56,69,82],
which
contrasts
with
the
fact
that
the
protein
has
been
detected
by
IF
in
spermatocytes
[86].
There
is
discrepancy
about
the
timing
when
AID
is
expressed
in
primordial
germ
cells
although
all
reports
find
mRNA
levels
sub-
stantially
lower
than
in
B
cells
(5–10%)
[56,84,85].
Similarly
AID
mRNA
levels
in
stem
cells
are
5–10%
of
those
found
in
B
cells
[56,83].
Thus,
the
amount
of
AID
required
for
its
proposed
role
in
epigenetic
reprogramming
seems
to
be
considerably
lower
than
that
required
for
antibody
diversification.
Even
lower
AID
mRNA
levels
than
in
stem
cells
have
been
measured
during
B
cell
develop-
ment
[65,66],
so
low
that
it
cannot
be
detected
by
using
an
AID-GFP
reporter
gene
[69].
And
yet,
convincing
evidence
suggests
that
it
is
enough
to
play
a
role
in
B
cell
tolerance
[65,66].
The
possibility
of
quite
low
levels
of
AID,
which
would
be
ineffective
for
antibody
A.
Orthwein,
J.M.
Di
Noia
/
Seminars
in
Immunology
24 (2012) 246–
254 249
diversification
in
a
normal
immune
response,
being
functional
is
even
more
striking
when
one
considers
the
post-translational
reg-
ulatory
mechanisms
restraining
AID
in
activated
B
cells.
3.2.
Post-translational
regulation
of
AID
in
B
cells
3.2.1.
AID
subcellular
localization
AID
is
predominantly
cytoplasmic
(see
Häsler,
Rada,
and
Neu-
berger,
this
issue)
in
steady-state
as
first
reported
for
AID-GFP
in
Ramos
B
cells
[87],
and
confirmed
by
subcellular
fractionation
and
IHC
in
primary
B
cells
[40,71,88].
The
deletion
or
mutation
of
the
C-terminal
10
amino
acids
of
AID
led
to
its
nuclear
accumu-
lation,
identifying
this
region
as
a
nuclear
export
signal
(NES)
and
demonstrating
that
AID
is
a
nuclear-cytoplasmic
shuttling
protein
[89–91].
This
leucine-rich
NES
is
a
typical
recognition
motif
for
the
exportin
CRM1,
as
was
demonstrated
by
using
the
CRM1-specific
inhibitor
leptomycin
B
[90,91]
and
by
coimmunoprecipitation
[92,93].
AID
gains
access
to
the
nucleus
by
active,
energy-dependent
transport
[93].
Due
to
its
small
size
(24
kDa),
AID
could
in
principle
diffuse
passively
through
the
nuclear
pores.
However,
it
is
actively
imported
into
the
nucleus,
as
demonstrated
by
its
capacity
to
con-
fer
nuclear
localization
to
large
proteins
that
are
well
above
the
nuclear
pore
passive
diffusion
cut-off
[93].
The
nuclear
localization
signal
(NLS)
of
AID
has
not
been
completely
defined.
AID
N-terminal
region
(roughly
residues
5–50)
contains
multiple
basic
residues,
a
characteristic
of
many
NLS,
and
it
is
clearly
a
major
part
of
it
[90,93].
However,
by
itself
it
is
not
sufficient
to
mediate
nuclear
import
of
heterologous
proteins,
which
requires
the
first
181
out
of
the
198
AID
amino
acids
[93].
There
are
other
residues
elsewhere
in
the
pro-
tein
that
form
part
of
and/or
are
critical
in
displaying
the
NLS,
which
prompted
the
suggestion
that
AID
has
a
conformational
NLS
[93].
The
factors
mediating
AID
nuclear
import
are
also
not
well
defined.
AID
binds
in
vitro
to
several
karyopherin-$
importins,
which
are
dedicated
nuclear
import
factors,
but
their
functionality
has
been
only
indirectly
inferred
because
oxidative
stress,
which
inhibits
karyopherin-$-mediated
nuclear
import
[94,95],
also
inhibits
AID
import
[93].
However,
AID
also
interacts
with
CTNNBL1,
a
nuclear
protein
presumed
to
work
in
mRNA
splicing,
which
binds
to
NLS
through
an
armadillo-like
domain
[96,97].
Nuclear
accumulation
of
an
AID
variant
with
a
mutated
NES
is
partially
compromised
in
DT40
CTNNBL1/
cells,
suggesting
a
role
for
CTNNBL1
in
AID
nuclear
import
[97].
Still,
CH12
CTNNBL1/
cells
did
not
show
any
reduction
in
CSR
demonstrating
that
it
is
at
least
redundant
with
another
mechanism
for
importing
AID
into
the
nucleus
[98].
Since
CTNNBL1
interacts
with
importin-$s
[97],
they
could
be
part
of
the
same
pathway.
GANP,
a
protein
associated
with
the
RNA
shut-
tling
machinery
[99],
has
also
been
proposed
to
mediate
AID
active
nuclear
import,
largely
based
on
the
observation
that
its
overex-
pression
increases
the
nuclear
fraction
of
AID
[100].
Regardless
of
the
pathway,
it
is
clear
that
AID
is
actively
imported
into
the
nucleus
while
its
small
size
would
allow
it
to
passively
diffuse,
which
raises
the
question
of
why
this
would
be
necessary.
The
explanation
could
be
that
there
is
a
need
to
counteract
cytoplasmic
retention,
which
prevents
its
diffusion
[93].
Indeed,
a
motif
in
AID
C-terminal
region,
overlapping
with,
but
being
distinct
from
the
NES,
is
able
to
limit
the
passive
diffusion
of
GFP
into
the
nucleus
[93].
Separation
of
function
mutations
exist
to
corroborate
that
these
two
sig-
nals
are
distinct
and
mediate
different
protein–protein
interactions
[93].
The
translation
factor
eEF1$
could
participate
in
retention,
at
least
in
part,
since
it
is
stoichiometrically
associated
with
AID
in
the
cytoplasm
and
mutating
the
proposed
AID
cytoplasmic
reten-
tion
motif
can
disrupt
this
interaction
[93,101].
However,
genetic
confirmation
has
not
been
possible
since
eEF1$is
an
essential
factor.
AID
subcellular
localization
is
dynamic
and
reflects
the
equi-
librium
between
nuclear
export
and
cytoplasmic
retention,
and
the
opposing
active
nuclear
import.
The
fraction
of
steady
state
nuclear
AID
resulting
from
this
equilibrium
has
been
consistently
estimated
to
be
around
<10%
[40,88,91].
Whether
this
equilibrium
is
regulated
(such
as
to
increase
diversification
after
some
signal-
ing
event,
at
a
particular
cell
cycle
stage
[102],
etc.),
or
whether
the
proportion
of
AID
in
the
nucleus
is
constant
at
all
times
[68,103],
is
unresolved.
Drastic
alterations
or
truncation
of
the
AID
NES
abolish
CSR,
most
likely
because
this
region
also
contains
a
domain
medi-
ating
interactions
required
for
CSR
[90,91,104,105].
Nevertheless,
the
relocalization
of
AID
into
the
nucleus
by
mutating
or
delet-
ing
the
NES
increases
SHM
and
immunoglobulin
gene
conversion
in
overexpression
studies
in
vitro
[90,104,105].
Similarly,
mutating
the
C-terminal
cytoplasmic
retention
signal
of
AID
leads
to
faster
nuclear
import
and
higher
SHM
and
CSR
[93,105].
The
interpretation
of
this
data
at
face
value,
indicates
that
alterations
in
the
mechanisms
of
AID
subcellular
localization
have
consequences
on
antibody
diversification
and
that
nuclear
exclu-
sion
limits
the
biological
activity
of
AID.
However,
there
are
caveats
in
this
interpretation
as
most
AID
nuclear
variants
tend
to
show
higher
catalytic
activity
[91,104,106]
and
lower
expression
lev-
els
[91,105,107].
The
contribution
of
these
characteristics
to
the
observed
effect
on
antibody
diversification
is
unknown
and
we
could
be
over
or
underestimating
the
magnitude
of
the
effect.
3.2.2.
AID
stability
The
stability
of
AID
is
directly
linked
to
its
subcellular
localiza-
tion
with
cytoplasmic
AID
being
much
more
stable
than
nuclear
AID
[107].
The
8
h
half-life
of
AID
in
B
cells
represents
in
fact
a
rough
average
of
its
2.5
h
nuclear
and
18–20
h
cytoplasmic
half-
life
[107].
Indeed,
inhibition
of
AID
nuclear
export
with
leptomycin
B
and
mutation
or
deletion
of
the
NES,
as
well
as
disruption
of
AID
cytoplasmic
retention,
increase
the
nuclear
fraction
of
AID
and
cor-
relate
with
a
decrease
in
its
half-life
[93,101,105,107,108].
On
the
other
hand,
restricting
AID
to
the
cytoplasm
resulted
in
increased
AID
half-life
[93,107].
There
are
some
mechanistic
explanations
for
this
difference
in
stability.
HSP90
interacts
with
AID
in
the
cytoplasm
and
prevents
its
polyubiquitination
and
therefore
its
proteasomal
degradation
[108].
The
first
step
in
the
HSP90
molec-
ular
chaperoning
pathway
and
stabilization
is
the
interaction
of
AID
with
the
HSP40
and
HSP70
system
[109].
In
particular,
there
is
a
specific
dependence
on
the
HSP40
DnaJa1,
since
its
deficiency
results
in
decreased
stability
and
reduced
levels
of
AID,
accompa-
nied
by
loss
of
biological
activity
[109].
Simultaneous
inhibition
of
HSP90
and
the
proteasome,
but
not
inhibiting
only
the
proteasome,
results
in
massive
accumulation
of
polyubiquitinated
AID
[108].
Thus,
there
seems
to
be
very
little
turnover
of
cytoplasmic
AID,
stabilization
being
the
default
pathway.
HSP90
is
critical
for
this
stabilization,
most
likely
just
after
AID
protein
synthesis,
but
does
not
exclude
the
possibility
that
later
complexes
with
eEF1$
[101]
or
importins
[93],
could
also
be
stabilizing.
It
has
also
been
proposed
that
some
maturation
step
such
as
post-translational
modification
or
oligomerization
could
stabilize
AID
[103,108].
Once
inside
the
nucleus,
AID
is
much
less
stable,
either
because
it
looses
stabilizing
interactions
and/or
it
is
actively
destabilized.
Unlike
cytoplasmic
AID,
nuclear
AID
seems
to
be
constantly
tar-
geted
to
the
proteasome
by
ubiquitin-dependent
and
-independent
pathways
[107,110].
Proteasomal
inhibition
is
sufficient
for
sub-
stantial
accumulation
of
polyubiquitinated
nuclear
AID
[107].
The
E3
ubiquitin
ligases
modifying
nuclear
AID
are
unknown.
MDM2
interacts
with
AID
and
could
be
one
of
them
[111].
However,
DT40
cells
deficient
in,
or
overexpressing,
MDM2
show
a
very
modest
increase
and
decrease
in
Ig
gene
conversion,
respectively
[111].
Thus,
MDM2
is
either
redundant
with
some
other
ubiquitin
lig-
ase
or
not
relevant
for
AID.
Interestingly,
AID
with
no
internal
250 A.
Orthwein,
J.M.
Di
Noia
/
Seminars
in
Immunology
24 (2012) 246–
254
lysine
residues
is
still
significantly
polyubiquitinated
[107],
sug-
gesting
either
N-terminal
ligation
or
a
non-canonical
pathway
of
polyubiquitination
for
AID.
A
proportion
of
AID
interacts
with,
and
is
targeted
for
degradation
by,
the
nuclear
protein
REG-%,
which
brings
proteins
for
proteasomal
degradation
in
an
ubiquitin-
and
ATP-independent
fashion
[110]
(see
below).
The
cytoplasmic
stabilization
by
the
DnaJa1–HSP90
pathway
is
critical
for,
and
limiting
in,
producing
AID.
Inhibiting
HSP90,
or
DnaJa1
farnesylation
(which
prevents
binding
to
AID),
as
well
as
DnaJa1
knockdown
or
knockout,
all
result
in
lower
endogenous
AID
levels
and
a
nearly
proportional
decrease
in
antibody
diversifica-
tion
by
SHM
and
CSR
[108,109].
This
data
is
in
keeping
with
the
results
from
AID
haploinsufficient
mice
[9,10];
supporting
the
view
that
AID
protein
is
limiting.
Also
consistent
with
this
view,
DnaJa1
overexpression
increases
AID
levels
and
antibody
diversification
in
vitro
[109].
Similarly,
REG-%deficiency
results
in
higher
AID
steady-state
levels,
increased
nuclear
AID
stability
and
CSR
in
mice
[110].
Thus,
modulating
AID
stability
seems
to
be
one
more
way
of
limiting
AID
expression
and
yet
another
mechanism
restraining
antibody
diversification.
In
the
absence
of
AID-specific
repressors,
HSP90
inhibitors
derived
from
geldanamycin,
which
are
already
in
clinical
trials,
or
farnesyltransferase
inhibitors
to
inactivate
DnaJa1,
could
be
exploited
to
indirectly
target
AID
[108,109].
They
offer
the
possi-
bility
to
treat
some
of
the
pathologies
associated
with
exacerbated
antibody
diversification,
like
autoimmune
diseases,
or
derived
from
AID
side
effects
such
as
imatinib
resistance
in
CML
or
tumor
clonal
evolution.
The
challenge
here
is
to
identify
a
disease
where
AID
implication
and
onset
can
be
predicted
to
use
these
inhibitors
at
the
right
time.
3.2.3.
AID
phosphorylation
In
addition
to
AID
protein
levels
and
subcellular
localization,
phosphorylation
regulates
AID
activity.
Five
evolutionary
con-
served
AID
residues,
have
been
found
to
be
phosphorylated
in
B
cells:
Ser3,
Thr27,
Ser38,
Thr140
and
Tyr184
[8,112–116].
Despite
Tyr184
being
very
close
to
the
NES
and
the
cytoplasmic
retention
signal,
and
Ser3,
Thr27
and
Ser38
being
close
to
or
immersed
in
the
NLS,
neither
phosphonull
nor
phosphomimicking
mutations
at
any
of
these
sites
seem
to
affect
AID
subcellular
localization
[103,115,117].
Also,
it
is
not
known
whether
phosphorylation
may
contribute
to
AID
stability.
Phosphorylation
of
Thr27
inhibits
AID-induced
deamination
and
CSR
in
vitro,
suggesting
it
modulates
AID
specific
activity
[112,113,117],
but
whether
this
is
used
in
vivo
to
regulate
AID
is
unknown.
Phosphorylation
at
Ser3
reduces
AID
biological
activity
ex
vivo
but
does
not
impact
its
catalytic
activity,
and
its
role
and
importance
in
vivo
are
unknown
[114].
On
the
other
hand,
phos-
phorylation
of
Ser38
and
Thr140
are
not
essential
for
AID
catalytic
activity
in
vitro,
but
both
significantly
increase
AID
biological
activ-
ity
in
vivo
[8,112,115,118,119].
Ser38
is
phosphorylated
by
PKA
on
chromatin,
which
allows
AID
to
interact
with
RPA
and
greatly
facil-
itates
CSR
and
SHM
[8,112,113,115,119,120].
However,
the
effect
of
mutating
Ser38
and
Thr140
in
vivo
was
much
more
pronounced
when
combined
with
AID
haploinsufficiency
[8,118],
thus
suggest-
ing
that
these
modifications
are
not
limiting
for
AID
activity.
4.
Conclusion
and
questions
4.1.
Regulation
of
AID
in
cells
with
high
versus
low
levels
of
AID
expression
The
post-translational
regulation
of
AID
(by
stabilization,
com-
partmentalization,
and
phosphorylation)
greatly
contributes
to
balancing
antibody
gene
diversification
and
widespread
DNA
damage
in
germinal
center
B
cells.
The
expected
(and
in
some
cases
known)
contribution
of
AID
to
cancer
is
by
the
same
mechanisms
it
uses
for
antibody
diversification:
induction
of
mutations
and
DNA
breaks.
Where
the
information
is
available,
the
AID
levels
in
lym-
phoid
malignancies
and
solid
tumors
from
human
patients
tend
to
be
within
the
same
range
as
those
in
B
cells.
In
many
of
these
cancers,
AID
post-translational
regulation
mechanisms
seem
to
be,
for
the
most
part,
functional.
Indeed,
many
B
cell
lymphoma
cell
lines
expressing
AID
have
been
used
to
study
compartmentaliza-
tion,
stability
and
phosphorylation
of
AID,
and
these
findings
were
then
confirmed
in
primary
cells.
In
almost
every
cell
line
tested,
AID
is
cytoplasmic
and
responds
to
leptomycin
B;
is
sensitive
to
HSP90
inhibition,
and
is
phosphorylated.
However,
there
could
be
quantitative
or
qualitative
differences
that
go
unnoticed
because
not
much
work
has
been
done
on
actual
activated
germinal
center
B
cells.
For
instance,
there
could
be
differences
in
the
relative
strength
of
the
different
mechanisms
regulating
AID
subcellular
localization
and/or
stability,
which
may
shift
the
balance
of
AID
regulation
and
promote
DNA
damage.
This
could
partly
explain
why
certain
tis-
sues
can
be
more
sensitive
to
transformation
than
B
cells
in
AID
transgenic
mice
(i.e.
thymus,
lung
or
testis)
[18,121].
Tissues
expressing
very
low
AID
levels
are
quite
a
different
cat-
egory
to
consider.
First
there
is
genetic
evidence
suggesting
that
the
expression
of
AID
during
early
B
cell
development
in
the
bone
marrow
plays
a
physiological
role
by
eliminating
autoreactive
B
cells
[65,66].
It
was
suggested
that
AID
expression
during
B
cell
development
mutates
the
IgV
region
[122,123].
This
is
reminis-
cent
of
what
is
observed
in
sheep,
rabbits
and
cattle
where
SHM
or
Ig
gene
conversion
contributes
to
the
diversity
of
the
primary
antibody
repertoire
[124–126].
In
humans,
AID
expression
during
early
B
cell
development
could
be
an
evolutionary
relic
or
may
con-
tribute
to
tolerance
by
mutating
autoreactive
clones,
thus
changing
their
specificity.
However,
given
the
almost
linear
proportionality
between
AID
protein
levels
and
SHM
in
germinal
center
B
cells,
it
is
difficult
to
imagine
that
the
1000-fold
lower
levels
of
AID
expressed
in
immature
B
cells
when
compared
to
germinal
center
B
cells
[66],
could
do
much
SHM.
The
mechanism
of
AID-mediated
B
cell
tol-
erance
is
unknown
but
one
wonders
whether
the
1000-fold
less
mRNA
linearly
translates
into
1000-fold
less
AID,
or
whether
only
10%
of
this
AID
is
in
the
nucleus
and
how
unstable
is
it.
Simi-
lar
questions
can
be
asked
about
cells
where
AID
is
proposed
to
underpin
epigenetic
reprogramming.
One
might
need
to
invoke
a
much
more
efficient
(and
error
free)
mechanism
for
AID-mediated
demethylation
than
for
SHM.
4.2.
A
network
or
a
cycle
for
AID
protein
regulation?
It
is
likely
that
multiple
mechanisms
were
selected
during
evo-
lution
to
restrain
the
activity
of
AID.
The
alternatives,
widespread
genomic
damage
or
immunodeficiency,
are
highly
deleterious
traits.
Nevertheless,
the
redundancy
of
mechanisms
with,
appar-
ently,
the
same
function
is
striking.
The
total
AID
levels
expressed
in
B
cells
are
much
higher
than
the
amount
of
it
that
can
be
found
in
the
nucleus
at
any
time,
and
there
are
multiple
mechanisms
to
enforce
this
nuclear
exclusion.
These
mechanisms
could
be
seen
as
a
cycle
or
as
a
network
(Fig.
1).
The
first
model
envisions
that
these
mechanisms
actually
form
a
circuit
(f.i.
release
from
HSP90
is
always
followed
by
AID
associated
to
cytoplasmic
retention,
which
precedes
nuclear
import,
etc.).
This
could
allow
mechanisms
with
apparently
the
same
effect
to
play
distinct
roles.
Cytoplasmic
reten-
tion
could
contribute
more
to
AID
nuclear
exclusion
than
nuclear
export
(as
the
largely
cytoplasmic
distribution
of
endogenous
AID
in
Ramos
cells
treated
with
leptomycin
B
could
suggest
[93]).
Nuclear
export
could
contribute
to
the
recycling
of
AID
from
the
nucleus
in
keeping
with
the
decay
observed
when
nuclear
export
is
inhibited.
An
alternative
model
could
be
that
there
are
several
A.
Orthwein,
J.M.
Di
Noia
/
Seminars
in
Immunology
24 (2012) 246–
254 251
AHA-1
Networ
kPharmacologi
cal d
est
abilizat
ion
Proteasome
Hsp90 inhibito
rs
FTI
HSC70 ATP
ADP
CYTOPLASMIC
RETENTION
CYTOPLASMIC
RETENTION
NUCLEAR
IMPORT
NUCLEAR
IMPORT
NUCLEAR
EXPORT
NUCLEAR
DEGRADATION
NUCLEAR
EXPORT
CYTOPLASMIC
DEGRADATION
NUCLEAR
DEGRADATION
TARGETING
TARGETING
STABILIZATION
STABILIZATION
Oligomerization ?
PTM ?
?
CRM1
DnaJa1
HSP90
AID eEF1α
CTNNBL1
GANP
?
Poly Ub REG-γ
Cycle
Proteasome
E3?
Transcription
RNA processing factors
PKA phosphorylation
AID
ATP
ADP
DnaJa1 HSP90
A
CB
E3 (?)
P
Importin
α
Fig.
1.
Major
post-translational
regulation
steps
affecting
AID
levels.
(A)
Schematic
representation
of
steps
participating
in
stability
and
subcellular
localization
regulation
represented
within
a
cycle.
Only
selected
AID
interacting
factors
are
shown.
AID
is
synthesized
in
the
cytoplasm,
where
unfolded
AID
is
met
by
the
HSP40-HSC70
system,
the
specific
action
of
the
HSP40
DnaJa1
allows
transferring
AID
into
the
HSP90
molecular
chaperoning
stabilization
cycle.
After
some
undefined
maturation
step,
or
conformational
change,
AID
is
passed
onto
eEF1$
and/or
other
cytoplasmic
retention
factors
before
active
nuclear
import.
A
number
of
factors
could
be
implicated,
alternatively
or
jointly
in
AID
nuclear
import.
Inside
the
nucleus
AID
is
either
exported
by
CRM1
or
targeted
to
the
Ig
loci
by
interacting
with
a
number
of
RNA
processing
factors,
where
it
is
phosphorylated
by
PKA.
AID
is
(subsequently)
degraded
in
the
nucleus
either
through
ubiquitin-
or
REG%-dependent
proteasomal
degradation.
(B)
Simplified
schematic
representation
of
the
same
steps
as
in
A,
but
in
the
form
of
a
network
in
which
most
pools
of
AID
are
interconnected
(see
text).
(C)
Schematic
representation
of
AID
cytoplasmic
degradation
following
inhibition
of
the
HSP90
molecular
chaperoning
pathway.
HSP90
inhibitors
prevent
the
ATP
hydrolysis
cycle
of
the
chaperone.
FTI,
farnesyltransferase
inhibitors,
prevent
farnesylation
of
DnaJa1,
which
is
required
for
binding
to
and
stabilization
of
AID.
Both
inhibitors
lead
to
polyubiquitination
and
proteasomal
degradation
of
cytoplasmic
AID.
options
after
each
regulatory
point,
with
as
many
possible
destina-
tions
competing
for
AID
(f.i.
release
from
HSP90
could
be
followed
by
cytoplasmic
retention
or
nuclear
import
or
degradation;
nuclear
AID
could
have
a
similar
probability
of
being
exported,
targeted
to
the
Ig
locus
or
degraded,
etc.).
Of
course,
a
third
possibility
would
be
a
mixture
of
both
models,
but
this
simplification
could
help
in
postulating
testable
hypothesis.
Much
experimental
work
is
still
needed.
The
germinal
center
is
probably
the
only
normal
tissue
in
which
the
presence
of
AID
pro-
tein
has
been
accurately
measured.
Detection
elsewhere
has
mostly
relied
on
RT-PCR
[56,67,82–84]
or
non-quantitative
immunode-
tection
by
IHC
or
IF.
Comparing
the
relative
levels
of
AID
protein
between
different
tissues;
developing
sensitive
methods
to
fol-
low
AID
in
precursor
B
or
stem
cells,
comparing
the
regulatory
252 A.
Orthwein,
J.M.
Di
Noia
/
Seminars
in
Immunology
24 (2012) 246–
254
mechanisms
of
AID
between
germinal
center
and
transformed
B
cells
or
non
B
cell
expression
sites,
would
all
contribute
to
under-
standing
AID.
Acknowledgments
We
thank
Stephen
Methot
for
reading
the
manuscript.
The
work
in
our
laboratory
was
funded
by
operating
grants
from
the
Canadian
Institutes
of
health
research
and
the
Cancer
Research
Society
and
an
infrastructure
grant
from
the
Canadian
Fund
for
Innovation,
leaders
opportunity
fund.
JMDN
is
supported
by
a
Canada
Research
Chair
Tier
2
and
AO
by
a
Cole
Foundation
doctoral
fellowship.
References
[1] Klein
U,
Dalla-Favera
R.
Germinal
centres:
role
in
B-cell
physiology
and
malig-
nancy.
Nature
Reviews
Immunology
2008;8:22–33.
[2]
Muramatsu
M,
Kinoshita
K,
Fagarasan
S,
Yamada
S,
Shinkai
Y,
Honjo
T.
Class
switch
recombination
and
hypermutation
require
activation-induced
cyti-
dine
deaminase
(AID),
a
potential
RNA
editing
enzyme.
Cell
2000;102:553–63.
[3]
Revy
P,
Muto
T,
Levy
Y,
Geissmann
F,
Plebani
A,
Sanal
O,
et
al.
Activation-
induced
cytidine
deaminase
(AID)
deficiency
causes
the
autosomal
recessive
form
of
the
hyper-IgM
syndrome
(HIGM2).
Cell
2000;102:565–75.
[4] Conticello
SG.
The
AID/APOBEC
family
of
nucleic
acid
mutators.
Genome
Biol-
ogy
2008;9:229.
[5]
Stavnezer
J,
Guikema
JE,
Schrader
CE.
Mechanism
and
regulation
of
class
switch
recombination.
Annual
Review
of
Immunology
2008;26:261–92.
[6] Peled
JU,
Kuang
FL,
Iglesias-Ussel
MD,
Roa
S,
Kalis
SL,
Goodman
MF,
et
al.
The
biochemistry
of
somatic
hypermutation.
Annual
Review
of
Immunology
2008;26:481–511.
[7]
Di
Noia
JM,
Neuberger
MS.
Molecular
mechanisms
of
antibody
somatic
hyper-
mutation.
Annual
Review
of
Biochemistry
2007;76:1–22.
[8] McBride
KM,
Gazumyan
A,
Woo
EM,
Schwickert
TA,
Chait
BT,
Nussenzweig
MC.
Regulation
of
class
switch
recombination
and
somatic
mutation
by
AID
phosphorylation.
The
Journal
of
Experimental
Medicine
2008;205:2585–94.
[9] Sernandez
IV,
de
Yebenes
VG,
Dorsett
Y,
Ramiro
AR.
Haploinsufficiency
of
activation-induced
deaminase
for
antibody
diversification
and
chromosome
translocations
both
in
vitro
and
in
vivo.
PLoS
One
2008;3:e3927.
[10]
Takizawa
M,
Tolarova
H,
Li
Z,
Dubois
W,
Lim
S,
Callen
E,
et
al.
AID
expression
levels
determine
the
extent
of
cMyc
oncogenic
translocations
and
the
inci-
dence
of
B
cell
tumor
development.
The
Journal
of
Experimental
Medicine
2008;205:1949–57.
[11]
Jiang
C,
Zhao
ML,
Diaz
M.
Activation-induced
deaminase
heterozygous
MRL/lpr
mice
are
delayed
in
the
production
of
high-affinity
pathogenic
antibodies
and
in
the
development
of
lupus
nephritis.
Immunology
2009;126:102–13.
[12]
Wu
X,
Stavnezer
J.
DNA
polymerase
beta
is
able
to
repair
breaks
in
switch
regions
and
plays
an
inhibitory
role
during
immunoglobulin
class
switch
recombination.
The
Journal
of
Experimental
Medicine
2007;204:1677–89.
[13]
Kaminski
DA,
Stavnezer
J.
Antibody
class
switching
differs
among
SJL,
C57BL/6
and
129
mice.
International
Immunology
2007;19:545–56.
[14] Durandy
A,
Peron
S,
Taubenheim
N,
Fischer
A.
Activation-induced
cytidine
deaminase:
structure–function
relationship
as
based
on
the
study
of
mutants.
Human
Mutation
2006;27:1185–91.
[15]
Rush
JS,
Liu
M,
Odegard
VH,
Unniraman
S,
Schatz
DG.
Expression
of
activation-induced
cytidine
deaminase
is
regulated
by
cell
division,
providing
a
mechanistic
basis
for
division-linked
class
switch
recombination.
Proceed-
ings
of
the
National
Academy
of
Sciences
of
the
United
States
of
America
2005;102:13242–7.
[16]
Muto
T,
Okazaki
IM,
Yamada
S,
Tanaka
Y,
Kinoshita
K,
Muramatsu
M,
et
al.
Negative
regulation
of
activation-induced
cytidine
deaminase
in
B
cells.
Pro-
ceedings
of
the
National
Academy
of
Sciences
of
the
United
States
of
America
2006;103:2752–7.
[17]
Robbiani
DF,
Bunting
S,
Feldhahn
N,
Bothmer
A,
Camps
J,
Deroubaix
S,
et
al.
AID
produces
DNA
double-strand
breaks
in
non-Ig
genes
and
mature
B
cell
lymphomas
with
reciprocal
chromosome
translocations.
Molecular
Cell
2009;36:631–41.
[18]
Okazaki
IM,
Hiai
H,
Kakazu
N,
Yamada
S,
Muramatsu
M,
Kinoshita
K,
et
al.
Constitutive
expression
of
AID
leads
to
tumorigenesis.
The
Journal
of
Experi-
mental
Medicine
2003;197:1173–81.
[19]
Teng
G,
Hakimpour
P,
Landgraf
P,
Rice
A,
Tuschl
T,
Casellas
R,
et
al.
MicroRNA-155
is
a
negative
regulator
of
activation-induced
cytidine
deami-
nase.
Immunity
2008;28:621–9.
[20]
Dorsett
Y,
McBride
KM,
Jankovic
M,
Gazumyan
A,
Thai
TH,
Robbiani
DF,
et
al.
MicroRNA-155
suppresses
activation-induced
cytidine
deaminase-mediated
Myc-Igh
translocation.
Immunity
2008;28:630–8.
[21] Bardwell
PD,
Woo
CJ,
Wei
K,
Li
Z,
Martin
A,
Sack
SZ,
et
al.
Altered
somatic
hypermutation
and
reduced
class-switch
recombination
in
exonuclease
1-
mutant
mice.
Nature
Immunology
2004;5:224–9.
[22]
Rada
C,
Di
Noia
JM,
Neuberger
MS.
Mismatch
recognition
and
uracil
excision
provide
complementary
paths
to
both
Ig
switching
and
the
A/T-focused
phase
of
somatic
mutation.
Molecular
Cell
2004;16:163–71.
[23]
Ehrenstein
MR,
Neuberger
MS.
Deficiency
in
Msh2
affects
the
efficiency
and
local
sequence
specificity
of
immunoglobulin
class-switch
recombination:
parallels
with
somatic
hypermutation.
EMBO
Journal
1999;18:3484–90.
[24]
Rada
C,
Williams
GT,
Nilsen
H,
Barnes
DE,
Lindahl
T,
Neuberger
MS.
Immunoglobulin
isotype
switching
is
inhibited
and
somatic
hyper-
mutation
perturbed
in
UNG-deficient
mice.
Current
Biology
2002;12:
1748–55.
[25]
Zaheen
A,
Boulianne
B,
Parsa
JY,
Ramachandran
S,
Gommerman
JL,
Martin
A.
AID
constrains
germinal
center
size
by
rendering
B
cells
susceptible
to
apoptosis.
Blood
2009;114:547–54.
[26] Ramiro
AR,
Jankovic
M,
Callen
E,
Difilippantonio
S,
Chen
HT,
McBride
KM,
et
al.
Role
of
genomic
instability
and
p53
in
AID-induced
c-myc-Igh
translocations.
Nature
2006;440:105–9.
[27]
Robbiani
DF,
Bothmer
A,
Callen
E,
Reina-San-Martin
B,
Dorsett
Y,
Difilippan-
tonio
S,
et
al.
AID
is
required
for
the
chromosomal
breaks
in
c-myc
that
lead
to
c-myc/IgH
translocations.
Cell
2008;135:1028–38.
[28] Dorsett
Y,
Robbiani
DF,
Jankovic
M,
Reina-San-Martin
B,
Eisenreich
TR,
Nussenzweig
MC.
A
role
for
AID
in
chromosome
translocations
between
c-myc
and
the
IgH
variable
region.
The
Journal
of
Experimental
Medicine
2007;204:2225–32.
[29]
Chiarle
R,
Zhang
Y,
Frock
RL,
Lewis
SM,
Molinie
B,
Ho
YJ,
et
al.
Genome-
wide
translocation
sequencing
reveals
mechanisms
of
chromosome
breaks
and
rearrangements
in
B
cells.
Cell
2011;147:107–19.
[30]
Klein
IA,
Resch
W,
Jankovic
M,
Oliveira
T,
Yamane
A,
Nakahashi
H,
et
al.
Translocation-capture
sequencing
reveals
the
extent
and
nature
of
chromo-
somal
rearrangements
in
B
lymphocytes.
Cell
2011;147:95–106.
[31]
Yamane
A,
Resch
W,
Kuo
N,
Kuchen
S,
Li
Z,
Sun
HW,
et
al.
Deep-sequencing
identification
of
the
genomic
targets
of
the
cytidine
deaminase
AID
and
its
cofactor
RPA
in
B
lymphocytes.
Nature
Immunology
2011;12:62–9.
[32]
Phan
RT,
Dalla-Favera
R.
The
BCL6
proto-oncogene
suppresses
p53
expression
in
germinal-centre
B
cells.
Nature
2004;432:635–9.
[33]
Liu
M,
Duke
JL,
Richter
DJ,
Vinuesa
CG,
Goodnow
CC,
Kleinstein
SH,
et
al.
Two
levels
of
protection
for
the
B
cell
genome
during
somatic
hypermutation.
Nature
2008;451:841–5.
[34]
Hasham
MG,
Donghia
NM,
Coffey
E,
Maynard
J,
Snow
KJ,
Ames
J,
et
al.
Widespread
genomic
breaks
generated
by
activation-induced
cytidine
deam-
inase
are
prevented
by
homologous
recombination.
Nature
Immunology
2010;11:820–6.
[35]
Pasqualucci
L,
Bhagat
G,
Jankovic
M,
Compagno
M,
Smith
P,
Muramatsu
M,
et
al.
AID
is
required
for
germinal
center-derived
lymphomagenesis.
Nature
Genetics
2008;40:108–12.
[36] Smit
LA,
Bende
RJ,
Aten
J,
Guikema
JE,
Aarts
WM,
van
Noesel
CJ.
Expression
of
activation-induced
cytidine
deaminase
is
confined
to
B-cell
non-Hodgkin’s
lymphomas
of
germinal-center
phenotype.
Cancer
Research
2003;63:3894–8.
[37] Lossos
IS,
Levy
R,
Alizadeh
AA.
AID
is
expressed
in
germinal
center
B-cell-like
and
activated
B-cell-like
diffuse
large-cell
lymphomas
and
is
not
correlated
with
intraclonal
heterogeneity.
Leukemia
2004;18:1775–9.
[38] Bodor
C,
Bognar
A,
Reiniger
L,
Szepesi
A,
Toth
E,
Kopper
L,
et
al.
Aber-
rant
somatic
hypermutation
and
expression
of
activation-induced
cytidine
deaminase
mRNA
in
mediastinal
large
B-cell
lymphoma.
British
Journal
of
Haematology
2005;129:373–6.
[39]
Deutsch
AJ,
Aigelsreiter
A,
Staber
PB,
Beham
A,
Linkesch
W,
Guelly
C,
et
al.
MALT
lymphoma
and
extranodal
diffuse
large
B-cell
lymphoma
are
targeted
by
aberrant
somatic
hypermutation.
Blood
2007;109:3500–4.
[40]
Pasqualucci
L,
Guglielmino
R,
Houldsworth
J,
Mohr
J,
Aoufouchi
S,
Polakiewicz
R,
et
al.
Expression
of
the
AID
protein
in
normal
and
neoplastic
B
cells.
Blood
2004;104:3318–25.
[41]
Jankovic
M,
Robbiani
DF,
Dorsett
Y,
Eisenreich
T,
Xu
Y,
Tarakhovsky
A,
et
al.
Role
of
the
translocation
partner
in
protection
against
AID-dependent
chro-
mosomal
translocations.
Proceedings
of
the
National
Academy
of
Sciences
of
the
United
States
of
America
2010;107:187–92.
[42]
Gauwerky
CE,
Huebner
K,
Isobe
M,
Nowell
PC,
Croce
CM.
Activation
of
MYC
in
a
masked
t(8;17)
translocation
results
in
an
aggressive
B-cell
leukemia.
Pro-
ceedings
of
the
National
Academy
of
Sciences
of
the
United
States
of
America
1989;86:8867–71.
[43]
Kovalchuk
AL,
duBois
W,
Mushinski
E,
McNeil
NE,
Hirt
C,
Qi
CF,
et
al.
AID-
deficient
Bcl-xL
transgenic
mice
develop
delayed
atypical
plasma
cell
tumors
with
unusual
Ig/Myc
chromosomal
rearrangements.
The
Journal
of
Experi-
mental
Medicine
2007;204:2989–3001.
[44]
Ramiro
AR,
Jankovic
M,
Eisenreich
T,
Difilippantonio
S,
Chen-Kiang
S,
Mura-
matsu
M,
et
al.
AID
is
required
for
c-myc/IgH
chromosome
translocations
in
vivo.
Cell
2004;118:431–8.
[45] Feldhahn
N,
Henke
N,
Melchior
K,
Duy
C,
Soh
BN,
Klein
F,
et
al.
Activation-
induced
cytidine
deaminase
acts
as
a
mutator
in
BCR-ABL1-transformed
acute
lymphoblastic
leukemia
cells.
The
Journal
of
Experimental
Medicine
2007;204:1157–66.
[46]
Gruber
TA,
Chang
MS,
Sposto
R,
Muschen
M.
Activation-induced
cytidine
deaminase
accelerates
clonal
evolution
in
BCR-ABL1-driven
B-cell
lineage
acute
lymphoblastic
leukemia.
Cancer
Research
2010;70:7411–20.
[47] Klemm
L,
Duy
C,
Iacobucci
I,
Kuchen
S,
von
Levetzow
G,
Feldhahn
N,
et
al.
The
B
cell
mutator
AID
promotes
B
lymphoid
blast
crisis
and
drug
resistance
in
chronic
myeloid
leukemia.
Cancer
Cell
2009;16:232–45.
A.
Orthwein,
J.M.
Di
Noia
/
Seminars
in
Immunology
24 (2012) 246–
254 253
[48]
McCarthy
H,
Wierda
WG,
Barron
LL,
Cromwell
CC,
Wang
J,
Coombes
KR,
et
al.
High
expression
of
activation-induced
cytidine
deaminase
(AID)
and
splice
variants
is
a
distinctive
feature
of
poor-prognosis
chronic
lymphocytic
leukemia.
Blood
2003;101:4903–8.
[49]
Palacios
F,
Moreno
P,
Morande
P,
Abreu
C,
Correa
A,
Porro
V,
et
al.
High
expression
of
AID
and
active
class
switch
recombination
might
account
for
a
more
aggressive
disease
in
unmutated
CLL
patients:
link
with
an
activated
microenvironment
in
CLL
disease.
Blood
2010;115:4488–96.
[50]
Babbage
G.
Immunoglobulin
heavy
chain
locus
events
and
expression
of
activation-induced
cytidine
deaminase
in
epithelial
breast
cancer
cell
lines.
Cancer
Research
2006;66:3996–4000.
[51] Lin
C,
Yang
L,
Tanasa
B,
Hutt
K,
Ju
BG,
Ohgi
K,
et
al.
Nuclear
receptor-induced
chromosomal
proximity
and
DNA
breaks
underlie
specific
translocations
in
cancer.
Cell
2009;139:1069–83.
[52]
Endo
Y,
Marusawa
H,
Kinoshita
K,
Morisawa
T,
Sakurai
T,
Okazaki
IM,
et
al.
Expression
of
activation-induced
cytidine
deaminase
in
human
hepatocytes
via
NF-kappaB
signaling.
Oncogene
2007;26:5587–95.
[53] Komori
J,
Marusawa
H,
Machimoto
T,
Endo
Y,
Kinoshita
K,
Kou
T,
et
al.
Activation-induced
cytidine
deaminase
links
bile
duct
inflammation
to
human
cholangiocarcinoma.
Hepatology
2008;47:888–96.
[54] Kou
T,
Marusawa
H,
Kinoshita
K,
Endo
Y,
Okazaki
I-m
Ueda
Y,
et
al.
Expres-
sion
of
activation-induced
cytidine
deaminase
in
human
hepatocytes
during
hepatocarcinogenesis.
International
Journal
of
Cancer
2006;120:469–76.
[55]
Shinmura
K,
Igarashi
H,
Goto
M,
Tao
H,
Yamada
H,
Matsuura
S,
et
al.
Aber-
rant
expression
and
mutation-inducing
activity
of
AID
in
human
lung
cancer.
Annals
of
Surgical
Oncology:
The
Official
Journal
of
the
Society
of
Surgical
Oncology
2011.
[56]
Morgan
HD,
Dean
W,
Coker
HA,
Reik
W,
Petersen-Mahrt
SK.
Activation-
induced
cytidine
deaminase
deaminates
5-methylcytosine
in
DNA
and
is
expressed
in
pluripotent
tissues:
implications
for
epigenetic
reprogramming.
Journal
of
Biological
Chemistry
2004;279:52353–60.
[57] Hsu
HC,
Yang
P,
Wu
Q,
Wang
JH,
Job
G,
Guentert
T,
et
al.
Inhibition
of
the
catalytic
function
of
activation-induced
cytidine
deaminase
promotes
apo-
ptosis
of
germinal
center
B
cells
in
BXD2
mice.
Arthritis
and
Rheumatism
2011;63:2038–48.
[58] Jiang
C,
Zhao
ML,
Scearce
RM,
Diaz
M.
Activation-induced
deaminase-
deficient
MRL/lpr
mice
secrete
high
levels
of
protective
antibodies
against
lupus
nephritis.
Arthritis
and
Rheumatism
2011;63:1086–96.
[59]
Jiang
C,
Foley
J,
Clayton
N,
Kissling
G,
Jokinen
M,
Herbert
R,
et
al.
Abrogation
of
lupus
nephritis
in
activation-induced
deaminase-deficient
MRL/lpr
mice.
Journal
of
Immunology
2007;178:7422–31.
[60]
Zan
H,
Zhang
J,
Ardeshna
S,
Xu
Z,
Park
SR,
Casali
P.
Lupus-prone
MRL/faslpr/lpr
mice
display
increased
AID
expression
and
extensive
DNA
lesions,
com-
prising
deletions
and
insertions,
in
the
immunoglobulin
locus:
concurrent
upregulation
of
somatic
hypermutation
and
class
switch
DNA
recombination.
Autoimmunity
2009;42:89–103.
[61]
White
CA,
Seth
Hawkins
J,
Pone
EJ,
Yu
ES,
Al-Qahtani
A,
Mai
T,
et
al.
AID
dysregulation
in
lupus-prone
MRL/Fas(lpr/lpr)
mice
increases
class
switch
DNA
recombination
and
promotes
interchromosomal
c-Myc/IgH
loci
translo-
cations:
modulation
by
HoxC4.
Autoimmunity
2011;44:585–98.
[62]
Xu
X,
Hsu
HC,
Chen
J,
Grizzle
WE,
Chatham
WW,
Stockard
CR,
et
al.
Increased
expression
of
activation-induced
cytidine
deaminase
is
associated
with
anti-
CCP
and
rheumatoid
factor
in
rheumatoid
arthritis.
Scandinavian
Journal
of
Immunology
2009;70:309–16.
[63]
Hase
K,
Takahashi
D,
Ebisawa
M,
Kawano
S,
Itoh
K,
Ohno
H.
Activation-induced
cytidine
deaminase
deficiency
causes
organ-specific
autoimmune
disease.
PLoS
One
2008;3:e3033.
[64]
Quartier
P,
Bustamante
J,
Sanal
O,
Plebani
A,
Debre
M,
Deville
A,
et
al.
Clinical,
immunologic
and
genetic
analysis
of
29
patients
with
autosomal
recessive
hyper-IgM
syndrome
due
to
activation-induced
cytidine
deaminase
defi-
ciency.
Clinical
immunology
(Orlando,
FL)
2004;110:22–9.
[65]
Meyers
G,
Ng
YS,
Bannock
JM,
Lavoie
A,
Walter
JE,
Notarangelo
LD,
et
al.
Activation-induced
cytidine
deaminase
(AID)
is
required
for
B-cell
tolerance
in
humans.
Proceedings
of
the
National
Academy
of
Sciences
of
the
United
States
of
America
2011;108:11554–9.
[66]
Kuraoka
M,
Holl
TM,
Liao
D,
Womble
M,
Cain
DW,
Reynolds
AE,
et
al.
Activation-induced
cytidine
deaminase
mediates
central
tolerance
in
B
cells.
Proceedings
of
the
National
Academy
of
Sciences
of
the
United
States
of
America
2011;108:11560–5.
[67]
Pauklin
S,
Sernandez
IV,
Bachmann
G,
Ramiro
AR,
Petersen-Mahrt
SK.
Estrogen
directly
activates
AID
transcription
and
function.
The
Journal
of
Experimental
Medicine
2009;206:99–111.
[68]
Ramiro
A,
Di
Noia
J.
Regulatory
mechanisms
of
AID
function.
In:
DNA
deami-
nation
and
the
immune
system.
Imperial
College
Press;
2010.
[69]
Crouch
EE,
Li
Z,
Takizawa
M,
Fichtner-Feigl
S,
Gourzi
P,
Montano
C,
et
al.
Regu-
lation
of
AID
expression
in
the
immune
response.
The
Journal
of
Experimental
Medicine
2007;204:1145–56.
[70]
Shaffer
AL,
Lin
KI,
Kuo
TC,
Yu
X,
Hurt
EM,
Rosenwald
A,
et
al.
Blimp-1
orches-
trates
plasma
cell
differentiation
by
extinguishing
the
mature
B
cell
gene
expression
program.
Immunity
2002;17:51–62.
[71]
Cattoretti
G,
Büttner
M,
Shaknovich
R,
Kremmer
E,
Alobeid
B,
Niedobitek
G.
Nuclear
and
cytoplasmic
AID
in
extrafollicular
and
germinal
center
B
cells.
Blood
2006;107:3967–75.
[72] Muramatsu
M,
Sankaranand
VS,
Anant
S,
Sugai
M,
Kinoshita
K,
Davidson
NO,
et
al.
Specific
expression
of
activation-induced
cytidine
deaminase
(AID),
a
novel
member
of
the
RNA-editing
deaminase
family
in
germinal
center
B
cells.
Journal
of
Biological
Chemistry
1999;274:18470–6.
[73]
Tran
TH,
Nakata
M,
Suzuki
K,
Begum
NA,
Shinkura
R,
Fagarasan
S,
et
al.
B
cell-specific
and
stimulation-responsive
enhancers
derepress
aicda
by
over-
coming
the
effects
of
silencers.
Nature
Immunology
2010;11:148–54.
[74]
Dedeoglu
F,
Horwitz
B,
Chaudhuri
J,
Alt
FW,
Geha
RS.
Induction
of
activation-induced
cytidine
deaminase
gene
expression
by
IL-4
and
CD40
ligation
is
dependent
on
STAT6
and
NFkappaB.
International
Immunology
2004;16:395–404.
[75] Machida
K,
Cheng
KT,
Sung
VM,
Shimodaira
S,
Lindsay
KL,
Levine
AM,
et
al.
Hepatitis
C
virus
induces
a
mutator
phenotype:
enhanced
mutations
of
immunoglobulin
and
protooncogenes.
Proceedings
of
the
National
Academy
of
Sciences
of
the
United
States
of
America
2004;101:4262–7.
[76] Matsumoto
Y,
Marusawa
H,
Kinoshita
K,
Endo
Y,
Kou
T,
Morisawa
T,
et
al.
Heli-
cobacter
pylori
infection
triggers
aberrant
expression
of
activation-induced
cytidine
deaminase
in
gastric
epithelium.
Nature
Medicine
2007;13:470–6.
[77]
Matsumoto
Y,
Marusawa
H,
Kinoshita
K,
Niwa
Y,
Sakai
Y,
Chiba
T.
Up-regulation
of
activation-induced
cytidine
deaminase
causes
genetic
aberrations
at
the
CDKN2b-CDKN2a
in
gastric
cancer.
Gastroenterology
2010;139:1984–94.
[78] Epeldegui
M,
Breen
EC,
Hung
YP,
Boscardin
WJ,
Detels
R,
Martinez-
Maza
O.
Elevated
expression
of
activation
induced
cytidine
deaminase
in
peripheral
blood
mononuclear
cells
precedes
AIDS-NHL
diagnosis.
AIDS
2007;21:2265–70.
[79]
Epeldegui
M,
Thapa
DR,
De
la
Cruz
J,
Kitchen
S,
Zack
JA,
Martinez-Maza
O.
CD40
ligand
(CD154)
incorporated
into
HIV
virions
induces
activation-
induced
cytidine
deaminase
(AID)
expression
in
human
B
lymphocytes.
PLoS
One
2010;5:e11448.
[80]
Gourzi
P,
Leonova
T,
Papavasiliou
FN.
A
role
for
activation-induced
cytidine
deaminase
in
the
host
response
against
a
transforming
retrovirus.
Immunity
2006;24:779–86.
[81] Gourzi
P,
Leonova
T,
Papavasiliou
FN.
Viral
induction
of
AID
is
independent
of
the
interferon
and
the
toll-like
receptor
signaling
pathways
but
requires
NF-kappaB.
The
Journal
of
Experimental
Medicine
2007;204:259–65.
[82]
MacDuff
DA,
Demorest
ZL,
Harris
RS.
AID
can
restrict
L1
retrotransposi-
tion
suggesting
a
dual
role
in
innate
and
adaptive
immunity.
Nucleic
Acids
Research
2009;37:1854–67.
[83]
Bhutani
N,
Brady
JJ,
Damian
M,
Sacco
A,
Corbel
SY,
Blau
HM.
Reprogramming
towards
pluripotency
requires
AID-dependent
DNA
demethylation.
Nature
2010;463:1042–7.
[84]
Popp
C,
Dean
W,
Feng
S,
Cokus
SJ,
Andrews
S,
Pellegrini
M,
et
al.
Genome-wide
erasure
of
DNA
methylation
in
mouse
primordial
germ
cells
is
affected
by
AID
deficiency.
Nature
2010;463:1101–5.
[85] Hajkova
P,
Jeffries
SJ,
Lee
C,
Miller
N,
Jackson
SP,
Surani
MA.
Genome-wide
reprogramming
in
the
mouse
germ
line
entails
the
base
excision
repair
path-
way.
Science
2010;329:78–82.
[86]
Schreck
S,
Buettner
M,
Kremmer
E,
Bogdan
M,
Herbst
H,
Niedobitek
G.
Activation-induced
cytidine
deaminase
(AID)
is
expressed
in
normal
sper-
matogenesis
but
only
infrequently
in
testicular
germ
cell
tumours.
The
Journal
of
Pathology
2006;210:26–31.
[87]
Rada
C,
Jarvis
JM,
Milstein
C.
AID-GFP
chimeric
protein
increases
hyper-
mutation
of
Ig
genes
with
no
evidence
of
nuclear
localization.
Proceedings
of
the
National
Academy
of
Sciences
of
the
United
States
of
America
2002;99:7003–8.
[88]
Schrader
CE,
Linehan
EK,
Mochegova
SN,
Woodland
RT,
Stavnezer
J.
Inducible
DNA
breaks
in
Ig
S
regions
are
dependent
on
AID
and
UNG.
The
Journal
of
Experimental
Medicine
2005;202:561–8.
[89]
Brar
SS,
Watson
M,
Diaz
M.
Activation-induced
cytosine
deaminase
(AID)
is
actively
exported
out
of
the
nucleus
but
retained
by
the
induction
of
DNA
breaks.
The
Journal
of
Biological
Chemistry
2004;279:26395–401.
[90]
Ito
S,
Nagaoka
H,
Shinkura
R,
Begum
NA,
Muramatsu
M,
Nakata
M,
et
al.
Activation-induced
cytidine
deaminase
shuttles
between
nucleus
and
cytoplasm
like
apolipoprotein
B
mRNA
editing
catalytic
polypeptide
1.
Pro-
ceedings
of
the
National
Academy
of
Sciences
of
the
United
States
of
America
2004;101:1975–80.
[91]
McBride
KM,
Barreto
V,
Ramiro
AR,
Stavropoulos
P,
Nussenzweig
MC.
Somatic
hypermutation
is
limited
by
CRM1-dependent
nuclear
export
of
activation-induced
deaminase.
The
Journal
of
Experimental
Medicine
2004;199:1235–44.
[92]
Ellyard
JI,
Benk
AS,
Taylor
B,
Rada
C,
Neuberger
MS.
The
dependence
of
Ig
class-
switching
on
the
nuclear
export
sequence
of
AID
likely
reflects
interaction
with
factors
additional
to
Crm1
exportin.
European
Journal
of
Immunology
2011;41:485–90.
[93]
Patenaude
AM,
Orthwein
A,
Hu
Y,
Campo
VA,
Kavli
B,
Buschiazzo
A,
et
al.
Active
nuclear
import
and
cytoplasmic
retention
of
activation-induced
deaminase.
Nature
Structural
&
Molecular
Biology
2009;16:517–27.
[94] Kodiha
M,
Chu
A,
Matusiewicz
N,
Stochaj
U.
Multiple
mechanisms
promote
the
inhibition
of
classical
nuclear
import
upon
exposure
to
severe
oxidative
stress.
Cell
Death
and
Differentiation
2004;11:862–74.
[95]
Miyamoto
Y,
Saiwaki
T,
Yamashita
J,
Yasuda
Y,
Kotera
I,
Shibata
S,
et
al.
Cellular
stresses
induce
the
nuclear
accumulation
of
importin
alpha
and
cause
a
con-
ventional
nuclear
import
block.
The
Journal
of
Cell
Biology
2004;165:617–23.
[96]
Conticello
SG,
Ganesh
K,
Xue
K,
Lu
M,
Rada
C,
Neuberger
MS.
Interaction
between
antibody-diversification
enzyme
AID
and
spliceosome-associated
factor
CTNNBL1.
Molecular
Cell
2008;31:474–84.
254 A.
Orthwein,
J.M.
Di
Noia
/
Seminars
in
Immunology
24 (2012) 246–
254
[97]
Ganesh
K,
Adam
S,
Taylor
B,
Simpson
P,
Rada
C,
Neuberger
MS.
CTNNBL1
Is
a
novel
nuclear
localization
sequence-binding
protein
that
recognizes
RNA-splicing
factors
CDC5L
and
Prp31.
The
Journal
of
Biological
Chemistry
2011;286:17091–102.
[98]
Han
L,
Masani
S,
Yu
K.
Cutting
edge:
CTNNBL1
is
dispensable
for
Ig
class
switch
recombination.
Journal
of
Immunology
2010;185:1379–81.
[99]
Wickramasinghe
VO,
McMurtrie
PI,
Mills
AD,
Takei
Y,
Penrhyn-Lowe
S,
Ama-
gase
Y,
et
al.
mRNA
export
from
mammalian
cell
nuclei
is
dependent
on
GANP.
Current
Biology
2010;20:25–31.
[100] Maeda
K,
Singh
SK,
Eda
K,
Kitabatake
M,
Pham
P,
Goodman
MF,
et
al.
GANP-
mediated
recruitment
of
activation-induced
cytidine
deaminase
to
cell
nuclei
and
to
immunoglobulin
variable
region
DNA.
The
Journal
of
Biological
Chem-
istry
2010;285:23945–53.
[101] Häsler
J,
Rada
C,
Neuberger
MS.
Cytoplasmic
activation-induced
cytidine
deaminase
(AID)
exists
in
stoichiometric
complex
with
translation
elonga-
tion
factor
1$(eEF1A).
Proceedings
of
the
National
Academy
of
Sciences
of
the
United
States
of
America
2011;108:18366–71.
[102] Ordinario
EC,
Yabuki
M,
Larson
RP,
Maizels
N.
Temporal
regulation
of
Ig
gene
diversification
revealed
by
single-cell
imaging.
Journal
of
Immunology
2009;183:4545–53.
[103] Patenaude
AM,
Di
Noia
JM.
The
mechanisms
regulating
the
subcellular
local-
ization
of
AID.
Nucleus
2010;1:325–31.
[104]
Barreto
VM,
Reina
San-Martin
BR,
Ramiro
AR,
McBride
KM,
Nussenzweig
MC.
C-terminal
deletion
of
AID
uncouples
class
switch
recombination
from
somatic
hypermutation
and
gene
conversion.
Molecular
Cell
2003;12:501–8.
[105]
Geisberger
R,
Rada
C,
Neuberger
MS.
The
stability
of
AID
and
its
function
in
class-switching
are
critically
sensitive
to
the
identity
of
its
nuclear-export
sequence.
Proceedings
of
the
National
Academy
of
Sciences
of
the
United
States
of
America
2009;106:6736–41.
[106]
Kohli
RM,
Abrams
SR,
Gajula
KS,
Maul
RW,
Gearhart
PJ,
Stivers
JT.
A
portable
hot
spot
recognition
loop
transfers
sequence
preferences
from
APOBEC
family
members
to
activation-induced
cytidine
deaminase.
The
Journal
of
Biological
Chemistry
2009;284:22898–904.
[107]
Aoufouchi
S,
Faili
A,
Zober
C,
D’Orlando
O,
Weller
S,
Weill
J-C,
et
al.
Pro-
teasomal
degradation
restricts
the
nuclear
lifespan
of
AID.
The
Journal
of
Experimental
Medicine
2008;205:1357–68.
[108]
Orthwein
A,
Patenaude
A-M,
Affar
EB,
Lamarre
A,
Young
JC,
Di
Noia
JM.
Regulation
of
activation-induced
deaminase
stability
and
antibody
gene
diversification
by
Hsp90.
The
Journal
of
Experimental
Medicine
2010;207:2751–65.
[109]
Orthwein
A,
Zahn
A,
Methot
SP,
Godin
D,
Conticello
SG,
Terada
K,
et
al.
Opti-
mal
functional
levels
of
activation-induced
deaminase
specifically
require
the
Hsp40
DnaJa1.
EMBO
Journal
2011.
[110] Uchimura
Y,
Barton
LF,
Rada
C,
Neuberger
MS.
REG-gamma
associates
with
and
modulates
the
abundance
of
nuclear
activation-induced
deaminase.
The
Journal
of
Experimental
Medicine
2011;208:2385–91.
[111]
MacDuff
DA,
Neuberger
MS,
Harris
RS.
MDM2
can
interact
with
the
C-
terminus
of
AID
but
it
is
inessential
for
antibody
diversification
in
DT40
B
cells.
Molecular
Immunology
2006;43:1099–108.
[112]
Basu
U,
Chaudhuri
J,
Alpert
C,
Dutt
S,
Ranganath
S,
Li
G,
et
al.
The
AID
anti-
body
diversification
enzyme
is
regulated
by
protein
kinase
A
phosphorylation.
Nature
2005;438:508–11.
[113]
Chatterji
M,
Unniraman
S,
McBride
KM,
Schatz
DG.
Role
of
activation-induced
deaminase
protein
kinase
A
phosphorylation
sites
in
Ig
gene
conversion
and
somatic
hypermutation.
Journal
of
Immunology
2007;179:5274–80.
[114]
Gazumyan
A,
Timachova
K,
Yuen
G,
Siden
E,
Di
Virgilio
M,
Woo
EM,
et
al.
Amino-terminal
phosphorylation
of
activation-induced
cytidine
deam-
inase
suppresses
c-myc/IgH
translocation.
Molecular
and
Cellular
Biology
2011;31:442–9.
[115]
McBride
KM,
Gazumyan
A,
Woo
EM,
Barreto
VM,
Robbiani
DF,
Chait
BT,
et
al.
Regulation
of
hypermutation
by
activation-induced
cytidine
deami-
nase
phosphorylation.
Proceedings
of
the
National
Academy
of
Sciences
of
the
United
States
of
America
2006;103:8798–803.
[116]
Pasqualucci
L,
Kitaura
Y,
Gu
H,
Dalla-Favera
R.
PKA-mediated
phosphorylation
regulates
the
function
of
activation-induced
deaminase
(AID)
in
B
cells.
Pro-
ceedings
of
the
National
Academy
of
Sciences
of
the
United
States
of
America
2006;103:395–400.
[117]
Demorest
ZL,
Li
M,
Harris
RS.
Phosphorylation
directly
regulates
the
intrin-
sic
DNA
cytidine
deaminase
activity
of
activation-induced
deaminase
and
APOBEC3G
protein.
The
Journal
of
Biological
Chemistry
2011;286:26568–75.
[118]
Cheng
HL,
Vuong
BQ,
Basu
U,
Franklin
A,
Schwer
B,
Astarita
J,
et
al.
Integrity
of
the
AID
serine-38
phosphorylation
site
is
critical
for
class
switch
recom-
bination
and
somatic
hypermutation
in
mice.
Proceedings
of
the
National
Academy
of
Sciences
of
the
United
States
of
America
2009;106:2717–22.
[119]
Vuong
BQ,
Lee
M,
Kabir
S,
Irimia
C,
Macchiarulo
S,
McKnight
GS,
et
al.
Specific
recruitment
of
protein
kinase
A
to
the
immunoglobulin
locus
regulates
class-
switch
recombination.
Nature
Immunology
2009;10:420–6.
[120]
Chaudhuri
J,
Khuong
C,
Alt
FW.
Replication
protein
A
interacts
with
AID
to
promote
deamination
of
somatic
hypermutation
targets.
Nature
2004;430:992–8.
[121] Takai
A,
Toyoshima
T,
Uemura
M,
Kitawaki
Y,
Marusawa
H,
Hiai
H,
et
al.
A
novel
mouse
model
of
hepatocarcinogenesis
triggered
by
AID
causing
deleterious
p53
mutations.
Oncogene
2009;28:469–78.
[122]
Han
JH,
Akira
S,
Calame
K,
Beutler
B,
Selsing
E,
Imanishi-Kari
T.
Class
switch
recombination
and
somatic
hypermutation
in
early
mouse
B
cells
are
medi-
ated
by
B
cell
and
toll-like
receptors.
Immunity
2007;27:64–75.
[123]
Mao
C,
Jiang
L,
Melo-Jorge
M,
Puthenveetil
M,
Zhang
X,
Carroll
MC,
et
al.
T
cell-independent
somatic
hypermutation
in
murine
B
cells
with
an
immature
phenotype.
Immunity
2004;20:133–44.
[124]
Lucier
MR,
Thompson
RE,
Waire
J,
Lin
AW,
Osborne
BA,
Goldsby
RA.
Mul-
tiple
sites
of
V
lambda
diversification
in
cattle.
Journal
of
Immunology
1998;161:5438–44.
[125] Reynaud
CA,
Garcia
C,
Hein
WR,
Weill
JC.
Hypermutation
generating
the
sheep
immunoglobulin
repertoire
is
an
antigen-independent
process.
Cell
1995;80:115–25.
[126]
Weinstein
PD,
Anderson
AO,
Mage
RG.
Rabbit
IgH
sequences
in
appendix
ger-
minal
centers:
VH
diversification
by
gene
conversion-like
and
hypermutation
mechanisms.
Immunity
1994;1:647–59.
... [14][15][16][17][18]). AID can cause genomic instability or chromosomal translocations [19][20][21][22][23] and the cytoplasmic localization of AID protects cells from deleterious consequences of nuclear AID activity. AID is sequestered in the cytoplasm [24] and requires active transport to enter the nucleus [25]. ...
... AID nuclear persistence is limited both by ubiquitin-dependent proteolysis [26,27] and export by the CRM1-dependent nuclear export pathway [15,16,28]. Nuclear AID is potentially toxic, and mutations that impair export or proteolysis of nuclear AID can compromise cell viability [22,27,29]. ...
Article
Full-text available
Author summary Activation-induced deaminase (AID) is a mutagenic factor that plays a critical role in immunoglobulin gene diversification and also functions in early development to reprogram methylated regions of the genome. AID must be tightly regulated to prevent compromising cell fitness, but how this occurs is not thoroughly understood. One level of regulation is known to be nuclear entry and exit, but spatiotemporal regulation of AID had not been examined at the level of single cells. We have carried out time-lapse imaging of individual cells expressing fluorescent-tagged AID. Our movies show that AID enters the nucleus in brief pulses, of about 30 minutes duration. Pulses occur in only a fraction of cells in the course of one cell cycle and appear to respond to a metronome intrinsic to individual cells. AID pulses are coordinated with pulses of P53, which regulates the cellular response to DNA damage. Pulsing may enable AID to synchronize both with factors that respond to AID-initiated damage and with factors that regulate transcription of AID target genes.
... This is already a mutagenic lesion causing a C:G to T:A base change after replication. Processing of the uracil by base-excision repair (BER) and mismatch repair (MMR) enzymes leads to the broader spectrum of point mutations characterizing SHM, and to DNA double strand breaks, which are necessary intermediates in CSR [ Figure 1 and reviewed in (14,15)]. ...
Article
Full-text available
The enzyme activation-induced cytidine deaminase (AID) initiates somatic hypermutation (SHM) and class switch recombination (CSR) of immunoglobulin (Ig) genes, critical actions for an effective adaptive immune response. However, in addition to the benefits generated by its physiological roles, AID is an etiological factor for the development of human and murine leukemias and lymphomas. This review highlights the pathological role of AID and the consequences of its actions on the development, progression, and therapeutic refractoriness of chronic lymphocytic leukemia (CLL) as a model disease for mature lymphoid malignancies. First, we summarize pertinent aspects of the expression and function of AID in normal B lymphocytes. Then, we assess putative causes for AID expression in leukemic cells emphasizing the role of an activated microenvironment. Thirdly, we discuss the role of AID in lymphomagenesis, in light of recent data obtained by NGS analyses on the genomic landscape of leukemia and lymphomas, concentrating on the frequency of AID signatures in these cancers and correlating previously described tumor-gene drivers with the presence of AID off-target mutations. Finally, we discuss how these changes could affect tumor suppressor and proto-oncogene targets and how they could be associated with disease progression. Collectively, we hope that these sections will help to better understand the complex paradox between the physiological role of AID in adaptive immunity and its potential causative activity in B-cell malignancies.
... Beyond transcriptional regulation of the AICDA gene, other levels of regulation have been proved to control AID activity. The compartmentalization of the protein, through a balance between the stabilization in the cytoplasm and the nucleus, ensures a mostly cytoplasmic presence of the enzyme and controls the extent of its effect on genomic DNA, while the NLS sequence allows for the active transport of AID to the nucleus Orthwein and Di Noia, 2012). Also, some micro-RNAs have been implicated in the stability of the AICDA transcript and are known to control the amount of AID that is expressed. ...
Thesis
Immunoglobulin (Ig) class switch recombination (CSR) takes place at the immunoglobulin heavy chain (IgH) constant locus upon B cell activation and results in a change of the isotype expressed. CSR is triggered by activation-induced cytidine deaminase (AID) and is dependent on inducible long-rangeenhancer/promoter looping and on germline transcription, which are controlled by the Eμ enhancer and the 3' regulatory region (3'RR) super-enhancer. Here, we characterize the role on switch transcription and recombination of g1E, a region located downstream of the Cg1 gene that bears marksof active enhancers and that interacts dynamically with both IgH enhancers upon B cell activation. We show that g1E deletion reduces CSR efficiency to IgA in CH12 cells and affects germline transcription and CSR in an isotype-specific manner in mice. On the other hand, whether transcription precedes orfollows looping to induce CSR is still unknown. To address this question, we targeted a transcriptional induction system based on the CRISPR/Cas9 technology to the Cg1 promoter in a background deficient for transcription and looping to study whether CSR could be restored.
... Target DNA transcription is required for SHM and is thought to provide the single-strand DNA template needed for AID to act (Keim et al., 2013;Pavri and Nussenzweig, 2011). The powerful mutagenic and genome-destabilizing potential of SHM suggests the need for careful regulation of the reaction, and in fact, AID is regulated at multiple levels, including tight control of Aicda transcription, posttranslational modification, protein degradation, an extensive protein interactome, and carefully orchestrated access of the enzyme to the nucleus (Keim et al., 2013;Orthwein and Di Noia, 2012). However, none of these AID-centric mechanisms explain how AID and SHM select specific regions of the genome on which to act. ...
Article
Full-text available
Somatic hypermutation (SHM) introduces point mutations into immunoglobulin (Ig) genes but also causes mutations in other parts of the genome. We have used lentiviral SHM reporter vectors to identify regions of the genome that are susceptible (“hot”) and resistant (“cold”) to SHM, revealing that SHM susceptibility and resistance are often properties of entire topologically associated domains (TADs). Comparison of hot and cold TADs reveals that while levels of transcription are equivalent, hot TADs are enriched for the cohesin loader NIPBL, super-enhancers, markers of paused/stalled RNA polymerase 2, and multiple important B cell transcription factors. We demonstrate that at least some hot TADs contain enhancers that possess SHM targeting activity and that insertion of a strong Ig SHM-targeting element into a cold TAD renders it hot. Our findings lead to a model for SHM susceptibility involving the cooperative action of cis-acting SHM targeting elements and the dynamic and architectural properties of TADs. : Senigl et al. show that genome susceptibility to somatic hypermutation (SHM) is confined within topologically associated domains (TADs) and is linked to markers of strong enhancers and stalled transcription and high levels of the cohesin loader NIPBL. Insertion of an ectopic SHM targeting element renders an entire TAD susceptible to SHM. Keywords: somatic hypermutation, activation induced deaminase, topologically associated domain, chromatin structure, chromatin loop extrusion, transcription factor
... While the molecular mechanisms triggering error-prone instead of error-free repair during SHM are largely elusive to date, mechanisms regulating AID activity are extensively studied and involve expression regulation via various transcription factors and miRNAs, balancing of cellular localization by cytosolic retention and nuclear import factors, as well as regulation of AID's nuclear stability and its targeting to Ig genes (16)(17)(18)(19). We have recently shown that PARP-1 is involved in AID regulation upon exogenous DNA damage, effectively leading to sequestration and stabilization of this mostly cytoplasmic enzyme in the cell nucleus (20). ...
Article
Full-text available
Affinity maturation of the humoral immune response depends on somatic hypermutation (SHM) of immunoglobulin (Ig) genes, which is initiated by targeted lesion introduction by activation-induced deaminase (AID), followed by error-prone DNA repair. Stringent regulation of this process is essential to prevent genetic instability, but no negative feedback control has been identified to date. Here we show that poly(ADP-ribose) polymerase-1 (PARP-1) is a key factor restricting AID activity during somatic hypermutation. Poly(ADP-ribose) (PAR) chains formed at DNA breaks trigger AID-PAR association, thus preventing excessive DNA damage induction at sites of AID action. Accordingly, AID activity and somatic hypermutation at the Ig variable region is decreased by PARP-1 activity. In addition, PARP-1 regulates DNA lesion processing by affecting strand biased A:T mutagenesis. Our study establishes a novel function of the ancestral genome maintenance factor PARP-1 as a critical local feedback regulator of both AID activity and DNA repair during Ig gene diversification.
... Target DNA transcription is required for SHM and is thought to provide the singlestrand DNA template needed for AID to act (Keim et al., 2013;Pavri and Nussenzweig, 2011). The powerful mutagenic and genome-destabilizing potential of SHM suggests the need for careful regulation of the reaction, and in fact, AID is regulated at multiple levels, including tight control of Aicda transcription, posttranslational modification, protein degradation, an extensive protein interactome, and carefully orchestrated access of the enzyme to the nucleus (Keim et al., 2013;Orthwein and Di Noia, 2012). However, none of these AID-centric mechanisms explain how AID and SHM select specific regions of the genome on which to act. ...
... AID R-mutants enter the nucleus but lack off-target activity. Nuclear access of AID is restricted, with~10% of AID being nuclear in homeostasis as a consequence of several mechanisms regulating AID nuclear residency and protein stability 33 . To exclude that the functional defect of the R-mutants was due to defective nuclear access, we analysed their subcellular localization. ...
Article
Full-text available
Activation-induced deaminase (AID) mutates the immunoglobulin (Ig) genes to initiate somatic hypermutation (SHM) and class switch recombination (CSR) in B cells, thus underpinning antibody responses. AID mutates a few hundred other loci, but most AID-occupied genes are spared. The mechanisms underlying productive deamination versus non-productive AID targeting are unclear. Here we show that three clustered arginine residues define a functional AID domain required for SHM, CSR, and off-target activity in B cells without affecting AID deaminase activity or Escherichia coli mutagenesis. Both wt AID and mutants with single amino acid replacements in this domain broadly associate with Spt5 and chromatin and occupy the promoter of AID target genes. However, mutant AID fails to occupy the corresponding gene bodies and loses association with transcription elongation factors. Thus AID mutagenic activity is determined not by locus occupancy but by a licensing mechanism, which couples AID to transcription elongation.
... Such activation events bring into AID's scope of action new set of genes harboring the transcriptional hallmarks defined by Álvarez-Prado et al. (2018). AID activation occurs in B cell lymphomas themselves, triggering mutation events from single base modifications up to "mutation storms" or kataegis, as well as in various can-cers and inflammatory processes, induced notably during viral or bacterial infections (Orthwein and Di Noia, 2012;Casellas et al., 2016). The prediction tools proposed in this paper will thus be of great help to further decipher AID off-targets outside the physiological germinal center immune response. ...
Article
Full-text available
In this issue of JEM, Álvarez-Prado et al. (https://doi.org/10.1084/jem.20171738) designed a DNA capture library allowing them to identify 275 genes targeted by AID in mouse germinal center B cells. Using the molecular features of these genes to feed a machine-learning algorithm, they determined that high-density RNA PolII and Spt5 binding—found in 2.3% of the genes—are the best predictors of AID specificity.
... AID can also target non-Ig genes and can lead to point mutations or chromosomal translocations such as IgH/Myc translocations. Due to these effects, AID is described as the first mutator enzyme (Hasler et al., 2012;Orthwein and Di Noia, 2012). The incidence of B-cell tumor development in animal models is much reduced under AID-deficient conditions (Nagaoka et al., 2010). ...
Article
Chronic lymphocytic leukemia (CLL) is the most common leukemia in Western countries. Cytogenetic lesions such as del13q14, del11q22.3, and del17p13 are identified in 50-60% of patients. Activation-induced cytidine deaminase (AID) plays a central role in somatic hyper mutation (SHM) and class switch recombination (CSR) and functions on Ig genes, but also target non-Ig genes, and over-expression of AID can lead to point mutations or translocations in non-Ig genes such as IgH/Myc translocation. Dicer and Drosha, which have a role in activation process of miRNA, also act in a double-strand DNA break (DSB) repair mechanism. In this study, whether the changes of AID, Dicer and Drosha expressions may be associated with both deletions and clinical outcomes in patients with CLL were investigated. AID expressions were increased in patients with CLL. However, cell lysate AID protein levels were only increased in patients with del17p or del11q who have poor prognosis. Decreased Dicer expressions were found in patients with deletion, whereas increased Drosha expressions were found in patients without deletion and with del13q. According to Rai and Binet staging systems, advanced-stage patients showed increased AID protein levels, decreased Dicer and Drosha expressions. Our findings may suggest that high AID expression and lower Dicer expression were observed in patients with CLL especially del17p and del11q and might associated with deletions such as del17p and del11q. AID, Dicer, and Drosha expressions might be used as an indicator of prognosis for CLL.
Chapter
Class switch recombination (CSR) generates isotype-switched antibodies with distinct effector functions essential for mediating effective humoral immunity. CSR is catalyzed by activation-induced deaminase (AID) that initiates DNA lesions in the evolutionarily conserved switch (S) regions at the immunoglobulin heavy chain (Igh) locus. AID-initiated DNA lesions are subsequently converted into DNA double stranded breaks (DSBs) in the S regions of Igh locus, repaired by non-homologous end-joining to effect CSR in mammalian B lymphocytes. While molecular mechanisms of CSR are well characterized, it remains less well understood how upstream signaling pathways regulate AID expression and CSR. B lymphocytes express multiple receptors including the B cell antigen receptor (BCR) and co-receptors (e.g., CD40). These receptors may share common signaling pathways or may use distinct signaling elements to regulate CSR. Here, we discuss how signals emanating from different receptors positively or negatively regulate AID expression and CSR.
Article
Full-text available
The immunological targets of estrogen at the molecular, humoral, and cellular level have been well documented, as has estrogen's role in establishing a gender bias in autoimmunity and cancer. During a healthy immune response, activation-induced deaminase (AID) deaminates cytosines at immunoglobulin (Ig) loci, initiating somatic hypermutation (SHM) and class switch recombination (CSR). Protein levels of nuclear AID are tightly controlled, as unregulated expression can lead to alterations in the immune response. Furthermore, hyperactivation of AID outside the immune system leads to oncogenesis. Here, we demonstrate that the estrogen–estrogen receptor complex binds to the AID promoter, enhancing AID messenger RNA expression, leading to a direct increase in AID protein production and alterations in SHM and CSR at the Ig locus. Enhanced translocations of the c-myc oncogene showed that the genotoxicity of estrogen via AID production was not limited to the Ig locus. Outside of the immune system (e.g., breast and ovaries), estrogen induced AID expression by >20-fold. The estrogen response was also partially conserved within the DNA deaminase family (APOBEC3B, -3F, and -3G), and could be inhibited by tamoxifen, an estrogen antagonist. We therefore suggest that estrogen-induced autoimmunity and oncogenesis may be derived through AID-dependent DNA instability.
Article
Full-text available
The cytidine deaminase AID hypermutates immunoglobulin genes but can also target oncogenes, leading to tumorigenesis. The extent of AID's promiscuity and its predilection for immunoglobulin genes are unknown. We report here that AID interacted broadly with promoter-proximal sequences associated with stalled polymerases and chromatin-activating marks. In contrast, genomic occupancy of replication protein A (RPA), an AID cofactor, was restricted to immunoglobulin genes. The recruitment of RPA to the immunoglobulin loci was facilitated by phosphorylation of AID at Ser38 and Thr140. We propose that stalled polymerases recruit AID, thereby resulting in low frequencies of hypermutation across the B cell genome. Efficient hypermutation and switch recombination required AID phosphorylation and correlated with recruitment of RPA. Our findings provide a rationale for the oncogenic role of AID in B cell malignancy.
Article
Full-text available
The enzyme activation-induced deaminase (AID) deaminates deoxycytidine at the immunoglobulin genes, thereby initiating antibody affinity maturation and isotype class switching during immune responses. In contrast, off-target DNA damage caused by AID is oncogenic. Central to balancing immunity and cancer is AID regulation, including the mechanisms determining AID protein levels. We describe a specific functional interaction between AID and the Hsp40 DnaJa1, which provides insight into the function of both proteins. Although both major cytoplasmic type I Hsp40s, DnaJa1 and DnaJa2, are induced upon B-cell activation and interact with AID in vitro, only DnaJa1 overexpression increases AID levels and biological activity in cell lines. Conversely, DnaJa1, but not DnaJa2, depletion reduces AID levels, stability and isotype switching. In vivo, DnaJa1-deficient mice display compromised response to immunization, AID protein and isotype switching levels being reduced by half. Moreover, DnaJa1 farnesylation is required to maintain, and farnesyltransferase inhibition reduces, AID protein levels in B cells. Thus, DnaJa1 is a limiting factor that plays a non-redundant role in the functional stabilization of AID.
Article
Full-text available
Activation-induced deaminase (AID) acts on the immunoglobulin loci in activated B lymphocytes to initiate antibody gene diversification. The abundance of AID in the nucleus appears tightly regulated, with most nuclear AID being either degraded or exported back to the cytoplasm. To gain insight into the mechanisms regulating nuclear AID, we screened for proteins interacting specifically with it. We found that REG-γ, a protein implicated in ubiquitin- and ATP-independent protein degradation, interacts in high stoichiometry with overexpressed nuclear AID as well as with endogenous AID in B cells. REG-γ deficiency results in increased AID accumulation and increased immunoglobulin class switching. A stable stoichiometric AID-REG-γ complex can be recapitulated in co-transformed bacteria, and REG-γ accelerates proteasomal degradation of AID in in vitro assays. Thus, REG-γ interacts, likely directly, with nuclear AID and modulates the abundance of this antibody-diversifying but potentially oncogenic enzyme.
Article
Ig repertoire diversification in cattle was studied in the ileal Peyer’s patch (IPP) follicles of young calves and in the spleens of late first-trimester bovine fetuses. To investigate follicular diversification, individual IPP follicles were isolated by microdissection; Vλ diversity was examined by RT-PCR and subsequent cloning and sequencing. When 52 intrafollicular sequences from a 4-wk-old calf were determined and compared, two major groups, one of 23 members and the other of 25, could be delineated. An examination of these groups revealed clear genealogic relationships that implicated in situ diversification of Vλ sequences within the confines of an IPP follicle. Vλ expression was also examined in early (95 and 110 gestational day) fetal bovine spleens. Although earlier studies in cattle and sheep implicated the IPP as a likely site of Ab diversification, a close investigation of Vλ sequences in late first-trimester fetal calves revealed that diversity appears in the early fetal spleen before the establishment of a diverse repertoire in the ileum. When the sequences for the fetal spleen were compared with an existing pool of germline sequences, we found evidence of possible gene conversion events and possible untemplated point mutations occurring in sequences recovered from fetal spleens. We conclude that IPP is not the sole site of Vλ diversification in cattle. Also, as suggested for rabbits, cattle may use both gene conversion and untemplated somatic point mutation to diversify their primary Vλ repertoire.
Chapter
A number of transcriptional, posttranscriptional and posttranslational mechanisms are known to regulate AID expression and function. Here we first overview the signaling pathways leading to AID expression in B cells in vitro, which mostly respond to surface receptors that include cytokine receptors such as IL4 and TGFβ, CD40 ligation or Toll-like receptors. Integration of these signals promotes the activation of Aicda (the gene encoding AID) transcription through various transcription factors, including the ubiquitous Sp1 or Sp3, lymphocyte-specific Pax5, E2A or HoxC4 as well as NFkB and hormone receptors. AID expression levels seem to be also subject to strict control, as evidenced by the functional impact of AID gene dose and AID overexpression in the various outcomes of its activity. At least two microRNAs (miR-155 and miR-181b) have been shown to contribute to this quantitative regulation. In addition, alternatively spliced variants of AID have been detected in malignant and healthy B cells, whose functional relevance remains unresolved. Finally, AID activity is exquisitely controlled by nuclear export, nuclear import, cytoplasmic retention and protein stability, which cooperatively restrict its nuclear concentration. The interconnections and relevance of this plethora of mechanisms remain a fascinating issue that is far from being understood. A number of transcriptional, posttranscriptional and posttranslational mechanisms are known to regulate AID expression and function. Here we first overview the signaling pathways leading to AID expression in B cells in vitro, which mostly respond to surface receptors that include cytokine receptors such as IL4 and TGFβ, CD40 ligation or Toll-like receptors. Integration of these signals promotes the activation of Aicda (the gene encoding AID) transcription through various transcription factors, including the ubiquitous Sp1 or Sp3, lymphocyte-specific Pax5, E2A or HoxC4 as well as NFkB and hormone receptors. AID expression levels seem to be also subject to strict control, as evidenced by the functional impact of AID gene dose and AID overexpression in the various outcomes of its activity. At least two microRNAs (miR-155 and miR-181b) have been shown to contribute to this quantitative regulation. In addition, alternatively spliced variants of AID have been detected in malignant and healthy B cells, whose functional relevance remains unresolved. Finally, AID activity is exquisitely controlled by nuclear export, nuclear import, cytoplasmic retention and protein stability, which cooperatively restrict its nuclear concentration. The interconnections and relevance of this plethora of mechanisms remain a fascinating issue that is far from being understood.
Article
Activation-induced cytidine deaminase (AID) is a B lymphocyte-specific DNA deaminase that acts on the Ig loci to trigger antibody gene diversification. Most AID, however, is retained in the cytoplasm and its nuclear abundance is carefully regulated because off-target action of AID leads to cancer. The nature of the cytosolic AID complex and the mechanisms regulating its release from the cytoplasm and import into the nucleus remain unknown. Here, we show that cytosolic AID in DT40 B cells is part of an 11S complex and, using an endogenously tagged AID protein to avoid overexpression artifacts, that it is bound in good stoichiometry to the translation elongation factor 1 alpha (eEF1A). The AID/eEF1A interaction is recapitulated in transfected cells and depends on the C-terminal domain of eEF1A (which is not responsible for GTP or tRNA binding). The eEF1A interaction is destroyed by mutations in AID that affect its cytosolic retention. These results suggest that eEF1A is a cytosolic retention factor for AID and extend on the multiple moonlighting functions of eEF1A.