ArticlePDF Available

Compensatory dendritic cell development mediated by BATF-IRF interactions

Authors:

Abstract and Figures

The AP1 transcription factor Batf3 is required for homeostatic development of CD8α(+) classical dendritic cells that prime CD8 T-cell responses against intracellular pathogens. Here we identify an alternative, Batf3-independent pathway in mice for CD8α(+) dendritic cell development operating during infection with intracellular pathogens and mediated by the cytokines interleukin (IL)-12 and interferon-γ. This alternative pathway results from molecular compensation for Batf3 provided by the related AP1 factors Batf, which also functions in T and B cells, and Batf2 induced by cytokines in response to infection. Reciprocally, physiological compensation between Batf and Batf3 also occurs in T cells for expression of IL-10 and CTLA4. Compensation among BATF factors is based on the shared capacity of their leucine zipper domains to interact with non-AP1 factors such as IRF4 and IRF8 to mediate cooperative gene activation. Conceivably, manipulating this alternative pathway of dendritic cell development could be of value in augmenting immune responses to vaccines.
Content may be subject to copyright.
Compensatory dendritic cell development mediated by BATF-
IRF interactions
Roxane Tussiwand1,§, Wan-Ling Lee1,§, Theresa L. Murphy1,§, Mona Mashayekhi1, KC
Wumesh1, Jörn C. Albring1, Ansuman T. Satpathy1, Jeffrey A. Rotondo1, Brian T. Edelson1,
Nicole M. Kretzer1, Xiaodi Wu1, Leslie A. Weiss2, Elke Glasmacher3, Peng Li4, Wei Liao4,
Michael Behnke2, Samuel S.K. Lam1, Cora T. Aurthur1, Warren J. Leonard4, Harinder
Singh3, Christina L. Stallings2, L. David Sibley2, Robert D. Schreiber1, and Kenneth M.
Murphy1,5,*
1Department of Pathology and Immunology, Washington University School of Medicine, 660 S.
Euclid Ave., St. Louis, MO 63110, USA
2Department of Molecular Microbiology, Washington University School of Medicine, St. Louis, MO
63110, USA
3Department of Discovery Immunology, Genentech, Inc., 1 DNA Way, S. San Francisco, CA,
94080
4Laboratory of Molecular Immunology and Immunology Center, National Heart, Lung, and Blood
Institute, National Institutes of Health, Bethesda, MD 20892-1674
5Howard Hughes Medical Institute, Washington University School of Medicine, 660 S. Euclid
Ave., St. Louis, MO 63110, USA
Abstract
The AP-1 transcription factor
Batf3
is required for homeostatic development of CD8α+ classical
dendritic cells that prime CD8 T-cell responses against intracellular pathogens. Here, we identify
an alternative,
Batf3
-independent pathway for their development operating during infection with
intracellular pathogens mediated by the cytokines IL-12 and IFN-γ. This alternative pathway
results from molecular compensation for
Batf3
provided by the related AP-1 factors
Batf
, which
also functions in T and B cells, and
Batf2
induced by cytokines in response to infection.
Reciprocally, physiologic compensation between
Batf
and
Batf3
also occurs in T cells for
expression of IL-10 and CTLA-4. Compensation among BATF factors is based on the shared
capacity of their leucine zipper domains to interact with non-AP-1 factors such as Irf4 and Irf8 to
mediate cooperative gene activation. Conceivably, manipulating this alternative pathway of
dendritic cell development could be of value in augmenting immune responses to vaccines.
*To whom correspondence should be addressed. Phone 314-362-2009, Fax 314-747-4888, kmurphy@wustl.edu.
§These authors contributed equally to this study.
Author Contributions R.T., W.L., T.L.M., and M.M. performed experiments with
Batf
−/−,
Batf2
−/−,
Batf3
−/− and DKO mice;
T.L.M. made and analyzed all Batf mutants; J.A.R. aided with EMSA; W.KC., J.C.A., and A.T.S. were involved with microarray
analysis and generation of mutant mice. J.C.A. was involved in generating of
Batf2
−/− mice. B.T.E. was involved with
L.
monocytogenes
analysis; N.M.K was involved with cross-presentation analysis. X.W. was involved with bioinformatic analysis;
L.A.W. C.L.S. were involved with Mtb infection; E.G. and H.S. helped identify EMSA probes; P.L., W.L. and W.J.L. performed
ChIP-Seq and helped identify EMSA probes; M.B. and D.S. aided with
T. gondii
infections; S.S.K.L., C.T.A. and R.D.S. aided with
tumor models. H.S. and W.J.L. provided helpful discussions. K.M.M. directed the work and wrote the manuscript. All authors
discussed the results and contributed to the manuscript.
The authors have no conflicting financial interests.
NIH Public Access
Author Manuscript
Nature
. Author manuscript; available in PMC 2013 April 25.
Published in final edited form as:
Nature
. 2012 October 25; 490(7421): 502–507. doi:10.1038/nature11531.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Batf1 and Batf3 are activator protein 1 (AP-1)2 transcription factors3,4 with immune-specific
functions5–8.
Batf
is required for development of T helper cells producing IL-17 (TH17) and
follicular helper T (TFH) cells5, and class-switch recombination (CSR) in B cells6,9.
Batf3
is
required for development of CD8α+ classical dendritic cells (cDCs) and related CD103+
DCs8 that cross-present antigens to CD8 T cells7 and produce IL-12 in response to
pathogens10.
We recently recognized a heterozygous phenotype for
Batf3
of 50% fewer
CX3CR1CD8α+ cDCs in
Batf3
+/− mice11, implying levels of Batf3 are limiting for CD8α+
DC development under homeostatic conditions. While analyzing
Toxoplasma gondii
infection in
Batf3
−/− mice, we observed evidence suggesting
Batf3
-independent CD8α+
cDC development10 based on apparent priming of pathogen specific CD8 T cells in
Batf3
−/−
mice treated with IL-12. Here, we report a
Batf3
-independent pathway of CD8α+ cDC
development that functions physiologically during infection by intracellular pathogens and
describe its molecular basis, which involves compensatory BATF factors.
Pathogens and IL-12 restore CD8α+ cDCs in Batf3/ mice
IL-12 administration to
Batf3
−/− mice before infection with type II Prugniaud (Pru)
T.
gondii
reversed their susceptibility by inducing IFN-γ production not only from NK cells
but also from CD8 T cells10, suggesting potentially restored cross-priming. To test this idea,
we infected
Batf3
−/− mice with the attenuated12 RHΔ
ku80Δrop
5 strain of
T. gondii
and
examined CD8α+ cDCs (Supplementary Fig. 1a). Surprisingly, CD8α+ cDCs reappeared in
spleens of
Batf3
−/− mice by day 10 after infection. Infection by
L. monocytogenes
also
restored CD8α+ cDCs in
Batf3
−/− mice (Supplementary Fig. 1b). Aerosolized infection with
Mycobacterium tuberculosis
(Mtb) caused a progressive restoration of CD8α+ cDCs in
Batf3
−/− mice and expanded CD8α+ cDCs in WT mice (Fig. 1a). Mtb infection of
Batf3
−/−
mice restored the missing lung-resident CD103+ cDCs (DEC205+ CD24+ CD4 Sirp-α
CD11b)8,13 (Supplementary Fig. 1c).
Batf3
−/− mice showed no difference in survival to
Mtb infection compared to WT mice (Supplemental Fig. 1d), suggesting the initial lack of
CD8α+ cDCs and peripheral CD103+ cDCs was not sufficient for lethality with Mtb. Since
IL-12 is important in control of Mtb14, we measured serum IL-12 in Mtb-infected WT and
Batf3
−/− mice (Fig. 1b). IL-12 was reduced in
Batf3
−/− mice in the first three weeks, but
increased after 6 weeks to approximately 50% of Mtb-infected WT mice.
We found that IL-12 administration restored CD8α+ cDCs in all backgrounds of
Batf3
−/−
mice (Fig. 1c, Supplementary Fig. 2a). Restored CD8α+ cDCs were functional for cross-
presentation7,15,16 (Fig. 2a, Supplementary Fig. 2b). Purified splenic CD8α+ and CD4+
cDCs from WT or
Batf3
−/− mice, treated with vehicle or IL-12, were tested for cross-
presentation. As a control, OT-I T cells proliferated in response to WT CD8α+ DCs, but not
CD4+ DCs, co-cultured with ovalbumin-loaded cells with or without IL-12 treatment (Fig.
2a). Importantly, OT-I T cells proliferated similarly in response to IL-12-induced
Batf3
−/−
CD8α+ DCs, but not
Batf3
−/− CD4+ DCs, as to WT CD8α+ cDCs. The CD8α+ cDCs
restored by IL-12 in
Batf3
−/− mice were more similar in gene expression to WT CD8α+ DCs
than to CD4+ cDCs (Supplementary Fig. 2c). 1855 genes differed more than 4-fold in
expression between IL-12-induced CD8α+ cDCs and CD4+ cDCs within
Batf3
−/− mice.
However, in comparing IL-12-induced CD8α+ cDCs in
Batf3
−/− mice with the rare CD8α+
cDCs from
Batf3
−/− C57BL/6 mice, or with CD8α+ cDCs in WT mice, there were only 34
and 206 genes respectively differing more than 4-fold in expression.
The IL-12-induced restoration of CD8α+ cDCs in
Batf3
−/− mice was dependent on IFN-γ
in
vivo
(Supplementary Fig. 2d). With administration of control antibody, IL-12 induced a 3-
fold increase in CD8α+ cDCs in WT mice and restored CD8α+ cDCs in
Batf3
−/− mice. Both
Tussiwand et al. Page 2
Nature
. Author manuscript; available in PMC 2013 April 25.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
effects were blocked by anti-IFN-γ antibody. IL-12 induced IFN-γ production from NK
cells but not T cells (Supplementary Fig. 2e), and IL-12 treatment was able to restore
CD8α+ cDCs in
Rag2
−/−
Batf3
−/− mice (Supplementary Fig. 2f), indicating that T cells and
B cells are not required for this effect.
Batf3
−/− mice fail to reject highly immunogenic H31m1 and D42m1 fibrosarcomas7.
However,
Batf3
−/− mice pre-treated with IL-12 either fully rejected or showed reduced
growth of these fibrosacromas (Fig. 2b, Supplementary Fig. 2g). Rejection was not simply
due to NK cell activation, since IL-12 treatment failed to alter tumor growth in
Rag2−/−
mice. Strikingly, IL-12-treated
Batf3
−/− mice re-established priming of CD8 T cells capable
of infiltrating tumors similar to WT mice (Fig. 2c). Mice with the inactivating IRF8 R249C
mutation17–19 failed to restore CD8α+ cDCs upon IL-12 administration (Supplementary Fig.
3a), indicating that restoration requires functional IRF8 protein. Further, these mice were
highly susceptible to infection by aerosolized Mtb20 (Supplementary Fig. 3b), suggesting
that
Batf3
−/− mice may resist Mtb infection by restoration of CD8α+ cDCs. Thus,
physiological
Batf3
-independent development of CD8α+ cDCs is induced by pathogens,
mediated by IL-12 and IFN-γ.
Batf, Batf2 and Batf3 cross-compensate in DCs and T cells
We asked if
Batf
could replace
Batf3
for
in vitro
cDC development7,21 (Fig. 3a). CD103+
Sirp-α cDCs do not develop in Flt3L-treated
Batf3
−/− bone marrow (BM) cultures.
Retroviral expression of
Batf3
into
Batf3
−/− BM progenitors restored CD103+ cDC
development. Notably,
Batf
fully and cell-intrinsically restored CD103+ Sirp-α cDC
development, while
c-Fos
was inactive (Supplementary Fig. 3c). CD103+ cDCs restored by
Batf
and
Batf3 in vitro
were functional, showing features of mature CD103+ cDCs,
including loss of Sirp-α and CD11b, upregulation of CD24, and selective production of
IL-12 in response to
T. gondii
antigen (Supplementary Fig. 3c-d). Reciprocally,
Batf3
but
not
c-Fos
restored cell-intrinsic IL-17a production by
Batf
−/− T cells and CSR in
Batf
−/− B
cells (Supplementary Fig. 3e–f). Thus,
Batf
and
Batf3
can molecularly cross-compensate for
several distinct lineage-specific functions, activities not shared by
c-Fos
.
We also identified
in vivo
compensation between
Batf
and
Batf3
in DCs. On the 129SvEv
and BALB/c backgrounds,
Batf3
−/− mice completely lack CD8α+ cDCs (Fig. 3b,
Supplementary Fig. 2a). In contrast, C57BL/6
Batf3
−/− mice retain a 3–7% population of
CD8α+ cDCs19,22 in spleen and unexpectedly retain a completely normal population of
CD8α+ cDCs in skin-draining inguinal lymph nodes (ILNs) (Fig. 3b). These CD8α+ cDCs
are functional, since C57BL/6
Batf3
−/− mice showed robust priming of CD8 T cells against
HSV infection in the footpad, while 129SvEv
Batf3
−/− mice, lacking CD8α+ cDCs in ILNs,
did not (Supplementary Fig. 3g). However, C57BL/6
Batf
−/−
Batf3
−/− (BATF1/3DKO) mice
lack CD8α+ DCs in ILNs and fail to prime T cells against HSV. Thus, retention of
functional CD8α+ cDCs in ILNs of C57BL/6
Batf3
−/− mice is
Batf
-dependent.
Batf
and
Batf3
also compensate in expression of genes by T cells. IL-4 and IL-10 production
were not substantially affected in either
Batf
−/− or
Batf3
−/− TH2 cells (Fig. 3c,
Supplementary Fig. 3i–k). However, IL-10 is reduced at least 8-fold in BATF1/3DKO TH2
cells. Also, the inhibitory receptor CTLA-4 is partially reduced in
Batf
−/− TH2 cells but
reduced 3-fold further in BATF1/3DKO TH2 cells. In contrast, IFN-γ production by TH1
cells is unaffected by loss of
Batf
,
Batf3
or both.
We asked if IL-12-induced restoration of CD8α+ cDCs in
Batf3
−/− mice was due to
compensation by
Batf
(Supplementary Fig. 3h). Restoration of splenic CD8α+ cDCs in
IL-12-treated BATF1/3DKO mice was reduced to 5% from 11% in IL-12-treated
Batf3
−/−
mice. While
Batf
appears responsible for roughly half of the IL-12-induced restoration of
Tussiwand et al. Page 3
Nature
. Author manuscript; available in PMC 2013 April 25.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
CD8α+ cDCs in
Batf3
−/− mice, some residual compensation remained in BATF1/3DKO
mice. We asked if a third factor might compensate for
Batf
and
Batf3. Batf2
(SARI)23 is
closely related to
Batf
and
Batf3
and is induced by LPS and IFN-γ in macrophages and
CD103+ DC populations (Supplementary Fig. 4a–c). We found that
Batf2
was induced by
IFN-γ in WT and
Batf3
−/− DCs derived from Flt3L-cultured BM (Supplementary Fig. 4d–e)
and
in vivo
by IL-12 in DCs in
Batf3
−/− mice (Supplementary Fig. 4f). Induction of
Batf2
by
IFN-γ in cDCs made it a potential candidate to mediate IFN-γ-dependent compensation for
Batf3
.
We made
Batf2
−/− mice by targeting exons 1 and 2, eliminating Batf2 protein expression
(Supplementary Fig. 5).
Batf2
−/− mice had normal development of NK, T and B cells, pDCs,
neutrophils, resting cDCs, and peritoneal, liver and lung macrophages (Supplementary Fig.
6a–c). However,
Batf2
−/− mice displayed significantly decreased survival after infection by
T. gondii
(Pru) (Fig. 4a), although parasite burden and serum cytokines were similar to WT
mice (Supplementary Fig. 7a–b). Notably,
Batf2
−/− mice showed significantly decreased
numbers of lung-resident CD103+CD11b DCs and CD103 CD11b macrophages after
infection (Fig. 4b, Supplementary Fig. 7c). These changes were specific, since there were no
differences in other myeloid subsets (Supplementary Fig. 7d-e) Thus,
Batf2
plays a role in
maintaining numbers of
Batf3
-dependent CD103+ DCs in the lung following infection with
T. gondii
.
We asked if
Batf2
could compensate for DC defects in
Batf3
−/− mice and for T and B cell
defects in
Batf
−/− mice (Fig. 4c, Supplementary Fig. 8a-b). Like
Batf
, retroviral expression
of
Batf2
restored development of CD103+Sirp-α DCs in Flt3L-treated
Batf3
−/− BM. Unlike
Batf
and
Batf3
, expression of
Batf2
did not restore TH17 development
Batf
−/− T cells and
only weakly restored CSR in
Batf
−/− B cells. Thus,
Batf2
selectively compensates for
Batf
and
Batf3
in cDCs but not in T or B cells. We next examined
in vivo
IL-12-induced
restoration of CD8α+ cDCs in
Batf2
−/−,
Batf3−/−
and
Batf2
−/−
Batf3
−/− (BATF2/3DKO)
mice (Fig. 4d, Supplementary Fig. 8c). IL-12 treatment restored CD8α+ cDCs from <1% to
>5% of total cDCs in
Batf3
−/− mice, but this was significantly reduced to ~2% in
BATF2/3DKO mice. This suggests that
Batf2
is responsible for roughly half of IL-12-
induced CD8α+ cDC restoration in
Batf3
−/− mice. We found similar results with
in vitro
CD103+ cDC development. While GM-CSF restored only CD103 and not DEC205
expression in Flt3L-treated
Batf3
−/− BM cultures19, adding IFN-γ with GM-CSF restored
DEC205+CD103+CD11b DCs (Supplementary Fig. 8d-e). We used this system to analyze
compensation between
Batf
,
Batf2
, and
Batf3
(Supplementary Fig. 8d–e). Relative to WT
BM, CD103+DEC205+CD11b cDCs were partially reduced in
Batf
−/−,
Batf2
−/− and
Batf3
−/− singly-deficient BM, but were reduced to an even greater extent in both
BATF1/3DKO and BATF2/3DKO BM relative to all single-deficient BM cultures.
Collectively, these results suggest that
Batf
and
Batf2
both act in the cytokine-dependent
rescue of CD8α+ cDC development in
Batf3
−/− mice.
The BATF LZ domain interacts with IRF4 and IRF8
To understand BATF cross-compensation, we analyzed chimeric proteins containing fused
domains of Batf, Batf2, and c-Fos (Supplementary Fig. 9). c-Fos inactivity could result from
inhibitory actions of N- and C-terminal domains missing in Batf and Batf3 (Fig. 5a).
However, removing these domains from c-Fos (Δ5c-FosΔ3) or fusing the Batf DBD with
the c-Fos LZ (BBRFFΔ3) failed to restore CD103+ cDC development (Fig. 5a), TH17
development or CSR activity (Supplementary Fig. 10a–b). However, fusing the Batf LZ
with the c-Fos DBD (Δ5FFQBB) fully restored CD103+ cDC development in Ftl3L-treated
Batf3
−/− BM cultures, fully restored CSR in
Batf
−/− B cells and exhibited 20% of WT
activity for restoring IL-17 production. Further, Batf with a C-terminal GFP (Batf-GFP) or
Tussiwand et al. Page 4
Nature
. Author manuscript; available in PMC 2013 April 25.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
c-Fos (BBBF) domain had >40% of WT Batf activity in all three assays (Supplementary Fig.
11). Thus, the LZ domain, not the DBD, determines BATF specificity.
Removing the Batf2 C-terminal domain (Batf2 DM) allowed for full restoration of CD103+
cDC development, but led to only partial restoration of CSR activity (~15%), and no
restoration of TH17 development (Supplementary Fig. 10c–h). Fusing the Batf DBD with
the Batf2 LZ (B1HB2 DM) allowed for approximately 50% of WT Batf activity for CSR
and TH17 development in
Batf
−/− B and T cells. Fusing the Batf2 DBD with Batf LZ
(B2QB1) allowed for 50% of WT Batf activity in CD103+ cDC development, but provided
less than 5% WT Batf activity in TH17 development. Thus, the Batf2 DBD restricts activity
in T and B cells, but its LZ functions for all BATF lineage-specific activities.
A recent study24 proposed that Irf4 and Batf may interact. Correspondence between the
phenotypes of
Batf
−/− with
Irf4
−/− mice, lacking TH17 development and CSR, and
Batf3
−/−
with
Irf8
−/− mice, lacking CD8α+ cDC development (Supplementary Fig. 12)5–7,25–28,
suggested that Batf and Batf3 may cooperate with both Irf4 and Irf8. Further, ChIP-Seq
analysis of Batf and Irf4 revealed coincident binding to composite AP-1 and IRF motifs (Li
et al., In Press; Glasmacher et al, In Press). By analogy with the Ets-IRF composite elements
(EICE)29, the AP-1-IRF composite elements are designated as AICE (Glasmacher et al., In
Press; Li et al., In Press). Recalling the c-Fos LZ interaction with NF-AT that mediates
cooperative binding to
Il2
regulatory regions30, we therefore asked if the Batf LZ interacted
with non-AP-1 factors, including Irf4.
Electrophoretic mobility shift assays (EMSA) demonstrated interactions between BATF and
both Irf4 and Irf8 (Fig. 5, Supplementary Figs. 13, 14). The Batf/Jun complex that formed
on an AP-1 consensus probe1,2 was unchanged by addition of Irf4 or Irf8. Its abundance was
increased by additional JunB (Supplementary Fig. 13a). However, using an AICE from the
CTLA-4 locus, a slower mobility complex formed with addition of either
Irf4
or
Irf8
, which
required the IRF consensus element (Fig. 5b, Supplementary Fig. 13b). This Batf/Jun/Irf4
complex required both Batf and IRF protein, and Irf4 was unable to bind the AICE1 probe
without Batf (Fig. 5b). In contrast, Irf4 was able to bind to and Ets/IRF consensus
elements31 (EICE) in the presence of PU.1 independently of Batf (Fig. 5b). The Batf/Jun/
Irf4 complex had similar mobility to the Fos/Jun complex in activated cells, but could be
distinguished by antibody supershifts (Supplementary Fig. 13c). Spatial constraints on the
Batf/Irf4 interaction are suggested by lack of Batf/Irf4 complex formation on the AICE2
probe with reversed orientation of the AP-1 and IRF elements (Supplementary 13a, c).
The ability of chimeric Batf proteins to interact with Irf4 correlated with their functional
activity (Fig. 5c). The active Δ5FFQBB interacted with Irf4, but the inactive Δ5c-FosΔ3
and BBRFFΔ3 did not (Fig. 5c), although all three bound an AP-1 consensus
(Supplementary Fig. 13d). A Batf/IRF complex also formed in B and T cells, requiring both
the IRF and AP-1 consensus elements (Fig. 5d, Supplementary Fig. 14c), and involved
contribution by endogenous Irf4 and Irf8 (Supplementary Fig. 14a, b, d). The Batf/IRF
complex was completely absent in
Batf
−/− B cells, but was partially retained in single-
deficient
Irf4
−/− or
Irf8
−/− B cell extracts (Supplementary Fig. 14a-b). Supershifts using anti-
Irf4 and anti-Irf8 antibodies in
Irf4
−/− or
Irf8
−/− B cell nuclear extracts showed that the
endogenous Batf/IRF complex is composed of a mixture of Irf4 and Irf8 (Supplementary
Fig. 14b). Similarly, supershifts of primary TH17 cells demonstrate an interaction between
Batf and both Irf4 and Irf8 (Supplementary Fig. 14d).
We identified Batf LZ mutations which abrogated both function and Irf4 interactions (Fig.
5e–f). Positions b, c and f of the c-Fos LZ face away from the parallel coiled-coil of the Jun
LZ and mediate interactions with NF-AT30 (Supplementary Fig. 9d). Several Batf mutations
Tussiwand et al. Page 5
Nature
. Author manuscript; available in PMC 2013 April 25.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
in these positions had no effect on activity (Supplementary Table 1, Supplementary Fig. 9c).
However, four residues together controlled nearly all BATF-specific activity (Fig. 5e,
Supplementary 15). Batf L56A, Batf K63D and Batf E77K showed no reduction in CD103+
cDC development, and retained >60% of WT Batf activity in CSR. In contrast, Batf H55Q
activity was reduced by nearly 70% (Fig. 5e). Double mutants Batf K63D E77K and Batf
H55Q L56A were reduced by 50% and 75%, but quadruple mutant Batf H55Q L56A K63D
E77K had less than 10% of WT Batf activity in CD103+ cDC development (Fig. 5e). This
loss of activity was specific since two other quadruple mutants, Batf E59Q D60Q K63D
E77K and Batf K63D R69Q K70D E77K, maintained >50% of WT Batf activity
(Supplementary Fig. 15c–d). A similar pattern was seen for CSR in
Batf
−/− B cells
(Supplementary Fig. 15b). The activity of these mutants correlated with Irf4 interaction (Fig.
5f). WT Batf and functional mutant K63D E77K, and both functional quadruple mutants
formed a complex with Irf4, whereas the less active Batf H55Q and Batf H55Q L56A, and
the completely inactive Batf H55Q L56A K63D E77K did not, although all were stably
expressed and could bind an AP-1 site (Supplementary 13e–f).
Discussion
We uncovered a cytokine-driven pathway for expansion of functional CD8α+ cDCs
occurring physiologically during infection by intracellular pathogens. This pathway relies on
functional compensation for
Batf3
provided by
Batf
and
Batf2
through a shared specificity
defined by the BATF LZ domain to support CD8α+ cDC and TH17 development, and CSR
in B cells. This compensatory pathway of CD8α+ cDCs development may provide a basis
for augmenting therapeutic immune responses. The basis of this shared specificity is the LZ
interaction with non-AP-1 factors, including Irf4 and Irf8, extending the repertoire of AP-1
interactions beyond the recognized association of Fos with NF-AT30. The ability of both
Batf and Batf3 to support both Irf4- and Irf8-dependent lineage activities implies that the
specificity for gene activation is determined largely by the IRF factors. An important next
step will be to determine how regulatory elements discriminate between these distinct
complexes.
Online Methods
Generation of Batf2/ mice
The targeting construct was assembled by Gateway recombination cloning system
(Invitrogen). To construct pENTR loxFRT rNEO, a floxed PGK–
neor
(phosphoglycerate
kinase promoter - neomycin phosphotransferase) gene cassette (1982 bp) was excised from
the pLNTK targeting vector34 using
Sal
I and
Xho
I. After incubation with Easy-A cloning
enzyme (Stratagene) and dNTPs to generate 3-A overhangs, the PGK–
neor
cassette was
ligated into the multiple cloning site of pGEM-T Easy (Promega). The resulting 2022 bp
PGK–
neor
gene cassette was released using
Not
I and ligated into the 2554 bp backbone of
Not
I digested pENTR lox-Puro35. Flippase Recognition Target (FRT) sites were
sequentially inserted at the
Sac
I and
Hind
III sites using DNA fragments generated by
following annealing oligonucleotides: SacII-FRT-A
(GAAGTTCCTATTCCGAAGTTCCTATTCTCTAGAAAGTATAGGAACTTCCGC) and
SacII-FRT-B
(GGAAGTTCCTATACTTTCTAGAGAATAGGAACTTCGGAATAGGAACTTCGC), or
HindIII-FRT-A
(AGCTTGAAGTTCCTATTCCGAAGTTCCTATTCTCTAGAAAGTATAGGAACTTC)
and HindIII-FRT-B
(AGCTGAAGTTCCTATACTTTCTAGAGAATAGGAACTTCGGAATAGGAACTTCA).
To construct pENTR-Batf2-5HA, the 5 homology arm was generated by PCR from
genomic DNA using the following oligonucleotides, which contain attB4 and attB1r site:
Tussiwand et al. Page 6
Nature
. Author manuscript; available in PMC 2013 April 25.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
GGGGACAACTTTGTATAGAAAAGTTGGCAGGCTGAAGCAGGGACAC, and
GGGGACTGCTTTTTTGTACAAACTTGCCCTAGCACACCCAGTTTCAGTTTC. The
attB4-attB1r PCR fragment was ligated into pDONR(P4-P1R) plasmid (Invitrogen) by BP
recombination reaction. To construct pENTR-Batf2-3HA, the 3 homology arm was
generated by PCR from genomic DNA using the following oligonucleotides which contain
attB2 and attB3site:
GGGGACAGCTTTCTTGTACAAAGTGGGCAGTCCCAGCATGACCCAGCCCCAGCA
TGA CCCAGCATCTGC, and
GGGGACAACTTTGTATAATAAAGTTGCCGTCTCACTCAGTTCTGTGTGTG. The
attB2-attB3 PCR fragment was ligated into pDONR (P2-P3) plasmid (Invitrogen) by BP
recombination reaction, following by the LR recombination reaction to generate final
targeting construct by using pENTR-Batf2-5HA, pENTR-Batf2-3HA, pENTR-loxFRT
rNEO, and pDEST DTA-MLS. The linearized vector was electroporated into EDJ22
embryonic stem cells, 129 SvEv background, and targeted clones were identified by
Southern blot analysis with 5 and 3 probes. Blastocyst injections were performed, and
male chimeras were bred to female 129S6/SvEv mice. Probes for Southern Blot analysis
were amplified from genomic DNA using the following primers: 5-
GTTGGTGTGAGGTGATGAGGTCCC -3 and 5-
CTTGACTTCCTAGACCAGGGGC-3 for the 5 probe; 5-
GGACCATATACTCTTACTGTTGAAACC-3 and 5-GGGCTGGGGACACACATG-3
for the 3 probe. Southern Blot analysis and genotyping PCR were performed to confirm
germline transmission and the genotype of the progeny. Primers used for genotyping PCR
are shown as follows: KO screen forward, 5-
GAAACTGAAACTGGGTGTGCTAGGG-3; KO screen reverse, 5
CGCCTTCTTGACGAGTTCTTCTGAG-3; WT screen reverse, 5-
GCCTTCCTTGTCTCTCTCCATAGCG-3.
Generation of Batf2-specific rabbit polyclonal antibody
Murine full-length Batf2 cDNA was cloned into pET-28a (+) expression vector (Novagen)
to generate recombinant Batf2 protein with His tagged. Recombinant Batf2 was then
expressed using
Escherichia coli
BL21 (Invitrogen). Purified recombinant Batf2 was used to
immunize New Zealand White rabbits (Harlan), and rabbit anti-mouse Batf2 sera were
collected and tested by ELISA and Western Blot analysis.
Pathogen infections
Listeria monocytogenes
and
T. gondii
infections were carried out as previously
described36,37. The type II avirulent Prugniaud (Pru) strain of
T. gondii
expressing a firefly
luciferase and GFP transgene (provided by J. Boothroyd, Stanford University, Palo Alto,
CA) was used for infections of
Batf2
−/− mice for the experiments shown in Fig. 4 and
Supplementary Fig. 7. Briefly, 800 tachyzoites were injected intraperitoneally (i.p.) into
mice. For Herpes simplex virus 1 (HSV-1) infections, mice were infected with the KOS
strain subcutaneously (s.c.) with 1.5 × 105 PFU/mouse in the foot pad. Viral supernatant and
HSV peptide were provided by M. Colonna. For
Mycobacterium tuberculosis
(
Mtb
)
infections, Erdman strain was grown to a density of 8×106 CFU/ml, and mice were exposed
to aerosol infection for 40 minutes using a Glas-Col aerosol exposure system (AES). After
24 hours, two mice per group were sacrificed, and lungs were harvested to determine
infection efficiency, which was about 100 CFU/lung. Lungs and spleens were collected at 1,
3 and 9 weeks post infection. The left lobe of the lung and one third of the spleen were used
for histological examination upon formalin fixation. The right lobe of the lung and two
thirds of the spleen were divided into CFU analysis and FACS analysis. For CFU analysis,
organs were homogenized and plated at various dilutions on 7H-10 culture plates, and after
three weeks, colonies were counted. For FACS analysis, organs were prepared as
Tussiwand et al. Page 7
Nature
. Author manuscript; available in PMC 2013 April 25.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
described38. For lungs, Ficoll gradient purification was performed prior to surface staining.
All samples were fixed upon staining for 20 minutes in 4% formalin. Serum was collected
by retro-orbital bleeding at week 1, 2, 3 and 6 and decontaminated by double filtering
through a 0.2mm filter.
Bioluminescence Imaging
Imaging was done as previously described39. In brief, mice were injected intraperitoneally
with D-luciferin (Biosynth AG, Switzerland) at 150 mg/kg and allowed to remain active for
5 minutes before being anesthetized with 2% isoflurane for 5 minutes. Animals were than
imaged with a Xenogen IVIS200 machine (Caliper Life Sciences). Data were analyzed with
the Living Image software (Caliper Life Sciences).
In vitro T cell restimulation after HSV infection
One week after infection, spleens were harvested and restimulated with the HSV peptide
HSV-gB2 (498-505) (Anaspec). Briefly, 2×106 splenocytes were restimulated for 5 hours in
the presence of Brefeldin A at 1 μg/mL and analyzed by FACS for intracellular IFN-γ and
TNF-α production as described later.
In vitro cross-presentation
DC cross-presentation of antigen to CD8+ OT-I T cells was assessed as previously
described40. Briefly, Spleens from naïve, vehicle or IL12-treated WT or
Batf3−/−
mice were
digested with collagenase B (Roche) and DNase I (Sigma-Aldrich). CD11c+ DCs were
obtained by negative selection using B220, Thy1.2, and DX5 microbeads followed by
positive selection with CD11c microbeads by MACS purification (Miltenyi Biotec). CD8α+
and CD4+ cDCs were cell sorted on a FACSAria II flow cytometer (post-sort purity >96%).
Splenocytes from Kb−/−Db−/−β2m−/− mice were prepared in serum-free medium, loaded with
10 mg/ml ovalbumin (EMD) by osmotic shock, and irradiated (13.5 Gy) as described
previously40. OT-I T cells were purified from OT-I/
Rag2−/−
mice by CD11c and DX5
negative selection, followed by positive selection with CD8α microbeads (purity >96%). T
cells were fluorescently labeled by incubation with 1 μM CFSE (Sigma-Aldrich) for 9 min
at 25°C at a density of 2 × 107 cells/ml. For the cross presentation assay, 5 × 104 –105
purified DCs were incubated with 5 × 104 –105 CFSE-labeled OT-I T cells in the presence
of increasing numbers of irradiated, ovalbumin-loaded Kb−/−Db−/−β2m−/− splenocytes (5 ×
103 to 1 × 105). After 3 days, OT-I T cell proliferation was analyzed by CSFE dilution
gating on CD3+ CD8+ CD45.1+ cells.
Tumor transplantation
MCA-induced fibrosarcomas were derived from 129/SvEv strain
Rag2−/−
or WT mice as
described previously41. Tumor cells were propagated
in vitro
and 106 tumor cells (129SvEv
fibrosarcoma tumor line H31m1 or D42m1) were injected s.c. in a volume of 150 μl
endotoxin-free PBS into the shaved flanks of naïve or IL12 conditioned (as described above)
wild type (WT)
Batf3
−/− and
Rag2
−/− recipient mice as described previously40. Injected
cells were >90% viable as assessed by trypan blue exclusion. Tumor size was measured on
the indicated days and is presented as the mean of two perpendicular diameters.
Tumor harvest
WT and naïve or IL12 treated
Batf3
−/− mice were inoculated with the H31m1 fibrosarcoma
tumor line. 11 days after, tumors were removed, minced and treated with 1μmg/ml type A
collagenase (Sigma) in HBSS (Hyclone) for 1h at 37°C. Cell suspension was analyzed by
FACS for CD8 T cell recruitment at the tumor site gating on CD45+, Thy1.2+ and CD8α+
cells.
Tussiwand et al. Page 8
Nature
. Author manuscript; available in PMC 2013 April 25.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Quantitative RT-PCR
For gene expression analysis, RNA was prepared from various cell types with either RNeasy
Mini Kit (Qiagen) or RNeasy Micro kit (QIAGEN), and cDNA was synthesized with
Superscript III reverse transcription (Invitrogen). Real-time PCR and a StepOnePlus Real-
Time PCR system (Applied Biosystems) were used according to the manufacturer’s
instructions, with the Quantitation Standard-Curve method and HotStart-IT SYBR Green
qPCR Master Mix (Affymetrix/USB). PCR conditions were 10 min at 95°C, followed by 40
two-step cycles consisting of 15 s at 95°C and 1 min at 60°C. Primers used for measurement
of
Batf2
expression are as follows: Batf2 qRT forward, 5-
GGCAGAAGCACACCAGTAAGG-3; Batf2 qRT reverse, 5-
GAAGGGCGTGGTTCTGTTTC-3. Hprt was used as normalization control, and primers
used are as follows: Hprt forward, 5-TCAGTCAACGGGGGACATAAA-3; and Hprt
reverse, 5-GGGGCTGTACTGCTTAACCAG-3.
DC preparation
Dendritic cells from lymphoid organs and non-lymphoid organs were harvested and
prepared as described 38. Briefly, spleens, MLNs, SLNs (inguinal), and liver were minced
and digested in 5 ml Iscove’s modified Dulbecco’s media + 10% FCS (cIMDM) with 250
μg/ml collagenase B (Roche) and 30 U/ml DNase I (Sigma-Aldrich) for 30 min at 37°C
with stirring. Cells were passed through a 70-μm strainer before red blood cells were lysed
with ACK lysis buffer. Cells were counted on a Vi-CELL analyzer, and 5–10 × 106 cells
were used per antibody staining reaction. Lung cell suspensions were prepared after
perfusion with 10 ml Dulbeccos PBS (DPBS) via injections into the right ventricle after
transection of the lower aorta. Dissected and minced lungs were digested in 5ml cIMDM
with 4mg/ml collagenase D (Roche) for 1 h at 37°C with stirring. Cells from the peritoneal
cavity were collected by washing the peritoneal cavity with 10 ml HBSS+2%FCS and 2mM
EDTA.
Bone marrow derived DC and macrophage
Bone marrow (BM)-derived DCs were generated from BM with 100 ng/mL of recombinant
Flt3L as described40. BM-derived macrophages were generated from BM with 20 ng/mL of
recombinant M-CSF (Peprotech) for 7 days, and rested in cIMDM without M-CSF for 24
hours before stimulation with various conditions as described in the figure legends.
Preparation of T. gondii lysate antigen (STAg)
T gondii
antigen was prepared by lysing tachyzoites of Pru strain in foreskin fibroblast
cultures through two cycles of freeze-thaw in liquid nitrogen. Lysate was then filtered,
aliquoted at a concentration of 1 mg/ml in PBS and stored at −80 degree. For DC
stimulation, 0.05 ug/ml were used and intracellular IL12 production was analyzed by FACS.
Cytokine induced CD8α+ DCs
WT and
Batf3
−/− mice bone marrow was prepared as described above. GM-CSF (10ng/ml)
and or IFN-γ (0.1ng/ml) was added to the cultures between day 8 and 10. Cells were
analysed two days after cytokine addition.
Antibodies and flow cytometry
Staining was performed at 4°C in the presence of Fc Block (anti CD16/32 clone 2.4G2,
BioXCell) in FACS buffer (DPBS + 0.5% BSA + 2 mm EDTA). The antibodies used for
DC analysis were as recently described42. Cells were analyzed on BD FACSCanto II or
FACSAria II and analyzed with FlowJo software (Tree star, Inc.).
Tussiwand et al. Page 9
Nature
. Author manuscript; available in PMC 2013 April 25.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Intracellular Cytokine Staining
For intracellular cytokine staining, cells were surface stained, then fixed in 2%
paraformaldehyde for 15 min at 4°C, permeabilized in DPBS + 0.1% BSA + 0.5% saponin,
and stained for intracellular cytokines as previously described43. Additional antibodies
included anti-TNF-α (MP6-XT22) and anti-IFN-γ (XMG1.2) from BioLegend.
ELISA and CBA
IL-12p40 concentration was measured from serum samples with the Mouse IL-12p40
OptEIA ELISA set (BD Bioscience) according to the manufacturer instructions. The
concentration of inflammatory cytokines was measured in the serum with the BD CBA
Mouse Inflammation Kit (BD Biosciences), and data were analyzed with FCAP Array
software (Soft Flow, Inc. USA).
Statistical analysis
Differences between groups in survival were analyzed by the log-rank test. Analysis of all
other data was done with an unpaired, two-tailed Student’s
t
test with a 95% confidence
interval (Prism; GraphPad Software, Inc.). P values less than 0.05 were considered
significant. *0.01 < p < 0.05, **0.001 < p < 0.01, ***p < 0.001, ****p < 0.0001.
Expression microarray analysis
Total RNA was isolated from cells using the Ambion RNAqueous-Micro Kit. For Mouse
Genome 430 2.0 Arrays, RNA was amplified, labeled, fragmented, and hybridized using the
3 IVT Express Kit (Affymetrix). Data were normalized and expression values were
modeled using DNA-Chip analyzer (dChip) software (www.dChip.org)44. Mouse Gene 1.0
ST Arrays, RNA was amplified with the WT Expression Kit (Ambion) and labeled,
fragmented, and hybridized with the WT Terminal Labeling and Hybridization Kit
(Affymetrix). Data was processed using RMA quantile normalization and expression values
were modeled using ArrayStar software (DNASTAR). All original microarray data have
been deposited in NCBI’s Gene Expression Omnibus and are accessible through GEO.
CD4+ T cell cultures
CD4+ T cells were purified using DynaBeads FlowComp mouse CD4 kit (Invitrogen) and
were activated on αCD3/αCD28 coated plates under the following culture conditions: TH1
conditions: αIL4 10μg/ml (11B11, BioXcell) 0.1μg/ml IFNγ (Peprotech), 10u/ml IL12, IL2
40u/ml; TH2 conditions: αIL12 10μg/ml (Tosh, BioXCell), αIFNγ 10μg/ml (XMG1.2,
BioXCell), 10ng/ml IL4 (Peprotech), IL2 40u/ml; TH17 conditiions: IL-6 25ng/ml
(Peprotech), TGFβ 2ng/ml (Peprotech), IL1β 10ng/ml (Peprotech), αIL4 10 μg/ml (11B11,
BioXCell), αIFNγ 10μg/ml (XMG1.2, BioXCell), and αIL12 10μg/ml (Tosh, BioXCell).
IL2 (40u/ml) was added for the TH17 cultures in Supplementary Figure 12a. Cells were
diluted three fold in fresh media on day 3. On day 5 cells were activated with PMA/
ionomycin for analysis of cytokines and surface markers by FACS. For some experiments,
cells were restimulated on day 7 under the same conditions and analyzed 5 days later.
Retroviral analysis of BATF functional activity
For CD8α+ DC development, BM was cultured with Flt3L as described40, infected with
retrovirus and 2μg/ml polybrene on day 1, and analyzed for development of cDCs (CD11c+
B220) on day 10.
CD4+ T cells were cultured under TH17 conditions43,45 and infected with retrovirus and
6μg/ml polybrene on day 1. Cells were analyzed on day 5 for cytokine expression by
Tussiwand et al. Page 10
Nature
. Author manuscript; available in PMC 2013 April 25.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
intracellular staining. For some experiments, cells were restimulated on day 7 under the
same conditions and analyzed 5 days later.
For analysis of class switch recombination, B220 cells purified by αB220 Macs microbeads
(Miltenyi) were cultured with LPS and IL4 as described46, infected with retrovirus and 6μg/
ml polybrene on day 1, and analyzed for switching to IgG1 on day 4 by FACS.
For infection of primary mouse cells, GFP-RV47 containing cDNAs for transcription factors,
chimeric proteins and mutated Batf were transfected into Phoenix E cells as described
previously48. Viral supernatants were collected 2 days later and concentrated by
centrifugation49. Cells were infected with viral supernatants by spin infection at 1800 rpm
for 45 min at room temperature.
Electromobility shift assays (EMSA)
For stable expression in 293FT cells, retroviruses (GFP-RV containing Batf, chimeric
proteins or Batf mutants, and/or truncated hCD4-RV47 containing Irf4 or Irf8 cDNA) were
packaged in Phoenix A cells and concentrated by centrifugation49. Infected 293FT cells
were sorted for retroviral marker expression and when indicated were transiently transfected
with JunB-GFP-RV using calcium phosphate. For transient expression, 293FT cells were
transfected with retroviral constructs using calcium phosphate. Nuclear extracts were
prepared 48hrs after transfection. For 293FT cells, nuclei were obtained after cellular lysis
with Buffer A containing 0.2% NP40 and nuclear extracts in Buffer C were dialyzed against
buffer D as described50. Nuclear extracts from T and B cells were prepared as described46.
EMSA was as described51 using 3μg of nuclear extract, 1mg poly dIdC (Sigma) and 7%T
3.3%C polyacrylamide gels. For competition assays, extracts were incubated with excess
unlabelled competitor DNA for 20 min before addition of labeled probe. For supershifts,
extracts were incubated with antibodies (αc-Fos (4) X (Santa Cruz) (for 293FT cells) or αc-
Fos 2G9C3 (Abcam) (for mouse T cells), αIrf4 (H- 140) X (Santa Cruz), αIrf8 (ICSBP)
C-19) X (Santa Cruz), αJun B (C-11) X (Santa Cruz) and rabbit αBatf 43, rabbit αBatf2
(described above), and rabbit αBatf3 (KC et al, submitted) for 1hr on ice prior to addition of
labeled probe and incubation at room temperature for 35minutes. The following pairs of
oligonucleotides were annealed to generate probes which were labeled with 32P-dCTP using
Klenow polymerase. The AP1consensus binding site is underlined. The Irf consensus
binding site is in italics. Mutated bases are in lower case.
AP1: 5GATCAGCTTCGCTTGATGAGTCAGCCGG/
5GATCCGGCTGACTCATCAAGCGAAG
AICE1 (from 3rd intron of mouse CTLA4):
5CTTGCCTTAGAGG
TTTC
GGGATGACTAATACTGTA/
5TCACGTACAGTATTAGTCATCCC
GAAA
CCTCTAAGG
AICE1 M1 (mutates the Irf consensus binding site):
5CTTGCCTTAGAGGccaCGGGATGACTAATACTGTA/
5TCACGTACAGTATTAGTCATCCCGtggCCTCTAAGG
AICE M2 (mutates the AP-1 consensus binding site):
5CTTGCCTTAGAGG
TTTC
GGGAgacCTAATACTGTA/
5TCACGTACAGTATTAGgtcTCCC
GAAA
CCTCTAAGG
Tussiwand et al. Page 11
Nature
. Author manuscript; available in PMC 2013 April 25.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
AICE2 (from 3rd intron of IL23R):
5GATGTTTCAGG
GAAA
GCACTGACTCACTGGCTCTCCA/
5GGTGGAGAGCCAGTGAGTCAGTGC
TTTC
CCTGAAA
EICE52: 5GAAAAAGAGAAATAAAAGGAAGT
GAAA
CCAAG/
5GATCCTTGG
TTTC
ACTTCCTTTTATTTCTCTTT
Eα: 5TCGACATTTTTCTGATTGGTTAAA/5GACTTTTAACCAATCAGAAAAATG
Plasmids
cDNAs for
Batf
,
Batf2
,
Batf3
,
Irf8
and
JunB
were generated by PCR from primary cells.
MigR1 containing HA tagged Irf4 was from H. Singh (Genentech). HA-Irf4 and Irf8 cDNAs
were subcloned into hCD4-RV47. Mutations in
Batf
were generated using QuickChange
mutagenesis with Batf -GFP-RV as the template. cDNAs for chimeric proteins were
generated by overlap extension and PCR and cloned into GFP-RV.
Supplementary Material
Refer to Web version on PubMed Central for supplementary material.
Acknowledgments
Supported by the Howard Hughes Medical Institute, National Institutes of Health (AI076427-02) and Department
of Defense (W81XWH-09-1-0185) (K.M.M.), the American Heart Association (12PRE8610005) (A.S.), German
Research Foundation (AL 1038/1-1) (J.C.A), American Society of Hematology Scholar Award and Burroughs
Welcome Fund Career Award for Medical Scientists (B.T.E.), and Cancer Research Institute predoctoral fellowship
(W.L.). We thank the ImmGen consortium32, Mike White for blastocyst injections and generation of mouse
chimeras, the Alvin J. Siteman Cancer Center at Washington University School of Medicine for use of the Center
for Biomedical Informatics and Multiplex Gene Analysis Genechip Core Facility. The Siteman Cancer Center is
supported in part by the NCI Cancer Center Support Grant P30 CA91842. IL-12 was a gift from Pfizer.
References
1. Dorsey MJ, et al. B-ATF: a novel human bZIP protein that associates with members of the AP-1
transcription factor family. Oncogene. 1995; 11:2255–2265. [PubMed: 8570175]
2. Wagner EF, Eferl R. Fos/AP-1 proteins in bone and the immune system. Immunol Rev. 2005;
208:126–140. [PubMed: 16313345]
3. Williams KL, et al. Characterization of murine BATF: a negative regulator of activator protein-1
activity in the thymus. Eur J Immunol. 2001; 31:1620–1627. [PubMed: 11466704]
4. Echlin DR, Tae HJ, Mitin N, Taparowsky EJ. B-ATF functions as a negative regulator of AP-1
mediated transcription and blocks cellular transformation by Ras and Fos. Oncogene. 2000;
19:1752–1763. [PubMed: 10777209]
5. Schraml BU, et al. The AP-1 transcription factor Batf controls T(H)17 differentiation. Nature. 2009;
460:405–409. [PubMed: 19578362]
6. Ise W, et al. The transcription factor BATF controls the global regulators of class-switch
recombination in both B cells and T cells. Nat Immunol. 2011
7. Hildner K, et al. Batf3 deficiency reveals a critical role for CD8alpha+ dendritic cells in cytotoxic T
cell immunity. Science. 2008; 322:1097–1100. [PubMed: 19008445]
8. Edelson BT, et al. Peripheral CD103+ dendritic cells form a unified subset developmentally related
to CD8alpha+ conventional dendritic cells. J Exp Med. 2010; 207:823–836. [PubMed: 20351058]
9. Betz BC, et al. Batf coordinates multiple aspects of B and T cell function required for normal
antibody responses. J Exp Med. 2010; 207:933–942. [PubMed: 20421391]
10. Mashayekhi M, et al. CD8a+ Dendritic Cells Are the Critical Source of Interleukin-12 that
Controls Acute Infection by Toxoplasma gondii Tachyzoites. Immunity. 2011; 35(2):249–259.
[PubMed: 21867928]
Tussiwand et al. Page 12
Nature
. Author manuscript; available in PMC 2013 April 25.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
11. Bar-On L, et al. CX3CR1+ CD8alpha+ dendritic cells are a steady-state population related to
plasmacytoid dendritic cells. Proc Natl Acad Sci U S A. 2010; 107:14745–14750. [PubMed:
20679228]
12. Behnke MS, et al. Virulence differences in Toxoplasma mediated by amplification of a family of
polymorphic pseudokinases. Proc Natl Acad Sci U S A. 2011; 108:9631–9636. [PubMed:
21586633]
13. Ginhoux F, et al. The origin and development of nonlymphoid tissue CD103+ DCs. J Exp Med.
2009; 206:3115–3130. [PubMed: 20008528]
14. Modlin RL, Barnes PF. IL12 and the human immune response to mycobacteria. Res Immunol.
1995; 146:526–531. [PubMed: 8839157]
15. den Haan JM, Lehar SM, Bevan MJ. CD8(+) but not CD8(−) dendritic cells cross-prime cytotoxic
T cells in vivo. J Exp Med. 2000; 192:1685–1696. [PubMed: 11120766]
16. Wilson NS, et al. Systemic activation of dendritic cells by Toll-like receptor ligands or malaria
infection impairs cross-presentation and antiviral immunity. Nat Immunol. 2006; 7:165–172.
[PubMed: 16415871]
17. Aliberti J, et al. Essential role for ICSBP in the in vivo development of murine CD8alpha +
dendritic cells. Blood. 2003; 101:305–310. [PubMed: 12393690]
18. Tailor P, Tamura T, Morse HC, Ozato K. The BXH2 mutation in IRF8 differentially impairs
dendritic cell subset development in the mouse. Blood. 2008; 111:1942–1945. [PubMed:
18055870]
19. Edelson BT, et al. Batf3-dependent CD11b(low/−) peripheral dendritic cells are GM-CSF-
independent and are not required for Th cell priming after subcutaneous immunization. PLoS
ONE. 2011; 6:e25660. [PubMed: 22065991]
20. Marquis JF, Lacourse R, Ryan L, North RJ, Gros P. Disseminated and rapidly fatal tuberculosis in
mice bearing a defective allele at IFN regulatory factor 8. J Immunol. 2009; 182:3008–3015.
[PubMed: 19234196]
21. Jackson JT, et al. Id2 expression delineates differential checkpoints in the genetic program of
CD8alpha(+) and CD103(+) dendritic cell lineages. EMBO J. 2011; 30:2690–704. [PubMed:
21587207]
22. Edelson BT, et al. CD8a+ Dendritic Cells Are an Obligate Cellular Entry Point for Productive
Infection by Listeria monocytogenes. Immunity. 2011; 35:236–248. [PubMed: 21867927]
23. Su ZZ, et al. Cloning and characterization of SARI (suppressor of AP-1, regulated by IFN). Proc
Natl Acad Sci U S A. 2008; 105:20906–20911. [PubMed: 19074269]
24. Ravasi T, et al. An atlas of combinatorial transcriptional regulation in mouse and man. Cell. 2010;
140:744–752. [PubMed: 20211142]
25. Brustle A, et al. The development of inflammatory T(H)-17 cells requires interferon-regulatory
factor 4. Nat Immunol. 2007; 8:958–966. [PubMed: 17676043]
26. Klein U, et al. Transcription factor IRF4 controls plasma cell differentiation and class-switch
recombination. Nat Immunol. 2006; 7:773–782. [PubMed: 16767092]
27. Tamura T, Ozato K. ICSBP/IRF-8: its regulatory roles in the development of myeloid cells. J
Interferon Cytokine Res. 2002; 22:145–152. [PubMed: 11846985]
28. Sciammas R, et al. Graded expression of interferon regulatory factor-4 coordinates isotype
switching with plasma cell differentiation. Immunity. 2006; 25:225–236. [PubMed: 16919487]
29. Brass AL, Zhu AQ, Singh H. Assembly requirements of PU. 1-Pip (IRF-4) activator complexes:
inhibiting function in vivo using fused dimers. EMBO J. 1999; 18:977–991. [PubMed: 10022840]
30. Chen L, Glover JN, Hogan PG, Rao A, Harrison SC. Structure of the DNA-binding domains from
NFAT, Fos and Jun bound specifically to DNA. Nature. 1998; 392:42–48. [PubMed: 9510247]
31. Eisenbeis CF, Singh H, Storb U. Pip, a novel IRF family member, is a lymphoid-specific, PU. 1-
dependent transcriptional activator. Genes Dev. 1995; 9:1377–1387. [PubMed: 7797077]
32. Heng TS, Painter MW. The Immunological Genome Project: networks of gene expression in
immune cells. Nat Immunol. 2008; 9:1091–1094. [PubMed: 18800157]
33. Ranganath S, et al. GATA-3-dependent enhancer activity in IL-4 gene regulation. J Immunol.
1998; 161:3822–3826. [PubMed: 9780146]
Tussiwand et al. Page 13
Nature
. Author manuscript; available in PMC 2013 April 25.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
34. Gorman JR, et al. The Ig(kappa) enhancer influences the ratio of Ig(kappa) versus Ig(lambda) B
lymphocytes. Immunity. 1996; 5:241–252. [PubMed: 8808679]
35. Iiizumi S, et al. Simple one-week method to construct gene-targeting vectors: application to
production of human knockout cell lines. Biotechniques. 2006; 41:311–316. [PubMed: 16989091]
36. Edelson BT, et al. CD8a+ Dendritic Cells Are an Obligate Cellular Entry Point for Productive
Infection by Listeria monocytogenes. Immunity. 2011; 35:236–248. [PubMed: 21867927]
37. Mashayekhi M, et al. CD8a+ Dendritic Cells Are the Critical Source of Interleukin-12 that
Controls Acute Infection by Toxoplasma gondii Tachyzoites. Immunity. 2011; 35:249–259.
[PubMed: 21867928]
38. Edelson BT, et al. Peripheral CD103+ dendritic cells form a unified subset developmentally related
to CD8alpha+ conventional dendritic cells. J Exp Med. 2010; 207:823–836. [PubMed: 20351058]
39. Saeij JP, Boyle JP, Grigg ME, Arrizabalaga G, Boothroyd JC. Bioluminescence imaging of
Toxoplasma gondii infection in living mice reveals dramatic differences between strains. Infect
Immun. 2005; 73:695–702. [PubMed: 15664907]
40. Hildner K, et al. Batf3 deficiency reveals a critical role for CD8alpha+ dendritic cells in cytotoxic
T cell immunity. Science. 2008; 322:1097–1100. [PubMed: 19008445]
41. Matsushita H, et al. Cancer exome analysis reveals a T-cell-dependent mechanism of cancer
immunoediting. Nature. 2012; 482:400–404. [PubMed: 22318521]
42. Satpathy AT, et al. Zbtb46 expression distinguishes classical dendritic cells and their committed
progenitors from other immune lineages. J Exp Med. 2012; 209:1135–1152. [PubMed: 22615127]
43. Schraml BU, et al. The AP-1 transcription factor Batf controls T(H)17 differentiation. Nature.
2009; 460:405–409. [PubMed: 19578362]
44. Li C, Wong WH. Model-based analysis of oligonucleotide arrays: expression index computation
and outlier detection. Proc Natl Acad Sci U S A. 2001; 98:31–36. [PubMed: 11134512]
45. Bending D, et al. Highly purified Th17 cells from BDC2.5NOD mice convert into Th1-like cells in
NOD/SCID recipient mice. J Clin Invest. 2009
46. Ise W, et al. The transcription factor BATF controls the global regulators of class-switch
recombination in both B cells and T cells. Nat Immunol. 2011; 12:536–543. [PubMed: 21572431]
47. Ranganath S, et al. GATA-3-dependent enhancer activity in IL-4 gene regulation. J Immunol.
1998; 161:3822–3826. [PubMed: 9780146]
48. Sedy JR, et al. B and T lymphocyte attenuator regulates T cell activation through interaction with
herpesvirus entry mediator. Nat Immunol. 2005; 6:90–98. [PubMed: 15568026]
49. Kanbe E, Zhang DE. A simple and quick method to concentrate MSCV retrovirus. Blood Cells
Mol Dis. 2004; 33:64–67. [PubMed: 15223013]
50. Dignam JD, Lebovitz RM, Roeder RG. Accurate transcription initiation by RNA polymerase II in
a soluble extract from isolated mammalian nuclei. Nucleic Acids Res. 1983; 11:1475–1489.
[PubMed: 6828386]
51. Szabo SJ, Gold JS, Murphy TL, Murphy KM. Identification of cis-acting regulatory elements
controlling interleukin-4 gene expression in T cells: roles for NF-Y and NF-ATc [published
erratum appears in Mol Cell Biol 1993 Sep;13(9):5928]. Mol Cell Biol. 1993; 13:4793–4805.
[PubMed: 8336717]
52. Eisenbeis CF, Singh H, Storb U. Pip, a novel IRF family member, is a lymphoid-specific, PU. 1-
dependent transcriptional activator. Genes Dev. 1995; 9:1377–1387. [PubMed: 7797077]
Tussiwand et al. Page 14
Nature
. Author manuscript; available in PMC 2013 April 25.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 1. Intracellular pathogens or IL-12 restore lymphoid CD8α+ cDCs and tissue-resident
CD103+ cDCs in Batf3−/− mice
a, Wild type (WT) and
Batf3
−/− (BATF3 KO) 129SvEv mice were uninfected (CTL) or
infected with Mtb, and spleens harvested and analyzed by FACS at the indicated time.
Histograms for indicated markers are gated as autofluorescentMHCIIhighCD11c+ cells.
Numbers are percent of cells in the gate. b, Serum IL-12 was measured from individual mice
(a) at the indicated time. c, Wild type (WT) and
Batf3
−/− (BATF3 KO) 129SvEv mice were
treated with vehicle (PBS) or IL-12 (IL12) and analyzed by FACS after 3 days as in (a).
Tussiwand et al. Page 15
Nature
. Author manuscript; available in PMC 2013 April 25.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 2. IL-12-induced CD8α+ cDCs in Batf3−/− mice can cross-present and mediate tumor
rejection
a, From mice in Fig. 1c, DCs were purified by sorting as CD3DX5MHCII+CD11c+Sirp-
αCD24+DEC205+ DCs (CD8DC) and CD3DX5MHCII+CD11c+Sirp-
α+CD24DEC205 DCs (CD4DC) and assayed for cross-presentation7. OT-I proliferation
in response to cDCs mixed with the indicated number of MHC class I-deficient ovalbumin
(Ova)-loaded splenocytes is shown. b, Wild type (WT) or
Batf3
−/− (BATF3 KO) mice
treated with vehicle (PBS) or with IL-12 (IL12) were inoculated with 1×106 H31m1
fibrosarcomas. Tumor size in individual mice is shown. c, Mice in (b) were analyzed by
FACS 11 days after H31m1 inoculation for CD8 T cell infiltration into tumors7.
Tussiwand et al. Page 16
Nature
. Author manuscript; available in PMC 2013 April 25.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 3. Batf compensates for CD8α+ cDC development in Batf3−/− mice
a, Wild type (WT), or
Batf3
−/− (BATF3KO) BM cells were infected with GFP-RV33
(Empty) or retrovirus expressing the indicated cDNA and cultured with Flt3L7. Histograms
for the indicated markers are for B220CD11c+ cells on day 10. Numbers are the percent of
cells in the gate. b, Inguinal lymph nodes from WT,
Batf3
−/− (BATF3 KO) or
Batf
−/−
Batf3
−/− mice (BATF1/3 DKO) on 129SvEv or C57BL/6 backgrounds were analyzed
by FACS. Shown are histograms for DEC205 and CD8α. c, CD4 T cells of the indicated
genotype were differentiated twice under TH2 conditions5 and analyzed by FACS for
intracellular IL-10.
Tussiwand et al. Page 17
Nature
. Author manuscript; available in PMC 2013 April 25.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 4. Batf2 compensates for Batf3 in CD8α+ and CD103+ cDC development during T. gondii
infection
a, Wild-type (WT) and
Batf2
−/− (BATF2KO) mice were infected with
T. gondii
and
monitored for survival. n=29 for WT (dashed line) and
Batf2
−/− (solid line) mice. b, Shown
are percentages of lung CD103+ DCs of total CD45.2+ cells for uninfected and infected (T.
gondii) mice on day 10. n=5 from one of three experiments. c, WT or
Batf3
−/− (BATF3KO)
BM cells were infected with the indicated retrovirus, cultured with Flt3L and analyzed by
FACS on day 10. d, Groups of 5 mice each of the indicated genotypes were treated with
vehicle (PBS) or IL-12 (IL12) and analyzed by FACS after 3 days. Shown are percentages
of CD8α+ cDCs as a total of splenic cDCs.
Tussiwand et al. Page 18
Nature
. Author manuscript; available in PMC 2013 April 25.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 5. BATF leucine zipper interactions with non-AP-1 factors mediate lineage-specific
actions
a, Structures of chimeric proteins are shown below a diagram of c-Fos. DNA binding
domain (DB), hinge (H), leucine zipper (LZ), amino- (5) and carboxy-terminus (3). Flt3L-
treated WT or
Batf3
−/− (BATF3 KO) BM infected with the indicated retrovirus were
analyzed after 10 days. b, 293FT cells expressing both
Batf
and
Irf4
(upper panel) or
Batf
and
Irf4
as indicated (lower panel) were analyzed by EMSA with the indicated probes and
antibodies c, 293FT cells expressing Irf4 (+) and the indicated Batf chimera were analyzed
by EMSA with the AICE1 probe. d, B cells were analyzed by EMSA with the indicated
probe and competitor oligonucleotides (comp). e,
Batf3
−/− (BATF3 KO) BM infected with
the indicated Batf retroviruses s encoding Batf were analyzed as in (a). f, 293FT cells
expressing the indicated Batf mutants were analyzed by EMSA with the AICE1 probe.
Tussiwand et al. Page 19
Nature
. Author manuscript; available in PMC 2013 April 25.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript

Supplementary resource (1)

... The protein panel investigated in this study comprised four pivotal proteins, namely BATF2, PTPRJ, RGS1 and VCAN, which potentially exert direct or indirect effects on pyroptosis and immune regulation within the TME across various cell populations. BATF2, a protein pivotal to the PRS calculation, plays a significant role in orchestrating the maturation of dendritic cells, T cells, B cells, and various immune cells and intricately participates in crucial biological processes and signaling pathways, including inflammatory responses, tumor immunity, and tumor cell proliferation and apoptosis [28][29][30][31][32]. The regulatory function of BATF2 in the progression of GC was extensively elucidated in our previous study [33]. ...
Article
Full-text available
Background: Pyroptosis plays a crucial role in immune responses. However, the effects of pyroptosis on tumor microenvironment remodeling and immunotherapy in gastric cancer (GC) remain unclear. Patients and Methods: Large-sample GEO data (GSE15459, GSE54129, and GSE62254) were used to explore the immunoregulatory roles of pyroptosis. TCGA cohort was used to elucidate multiple molecular events associated with pyroptosis, and a pyroptosis risk score (PRS) was constructed. The prognostic performance of the PRS was validated using postoperative GC samples from three public databases (n=925) and four independent Chinese medical cohorts (n=978). Single-cell sequencing and multiplex immunofluorescence were used to elucidate the immune cell infiltration landscape associated with PRS. Patients with GC who received neoadjuvant immunotherapy (n=48) and those with GC who received neoadjuvant chemotherapy (n=49) were enrolled to explore the value of PRS in neoadjuvant immunotherapy. Results: GC pyroptosis participates in immune activation in the tumor microenvironment and plays a powerful role in immune regulation. PRS, composed of four pyroptosis-related differentially expressed genes (BATF2, PTPRJ, RGS1, and VCAN), is a reliable and independent biomarker for GC. PRSlow is associated with an activated pyroptosis pathway and greater infiltration of anti-tumor immune cells, including more effector and CD4+ T cells, and with the polarization of tumor-associated macrophages in the tumor center. Importantly, PRSlow marks the effectiveness of neoadjuvant immunotherapy and enables screening of GC patients with combined positive score ≥1 who benefit from neoadjuvant immunotherapy. Conclusion: Our study demonstrated that pyroptosis activates immune processes in the tumor microenvironment. A low PRS correlates with enhanced infiltration of anti-tumor immune cells at the tumor site, increased pyroptotic activity, and improved patient outcomes. The constructed PRS can be used as an effective quantitative tool for pyroptosis analysis to guide more effective immunotherapeutic strategies for patients with GC.
... Cd74, Cd209a, Batf3, Tmem176b [19][20][21] (Fig. 2D, J) (Supplementary Tables S1 and S2). The monocyte subset expressing high mRNA levels of Irf7, Ifit3, Cxcl10 was termed the interferonstimulated gene (ISG) subset (Fig. 2D, Supplementary Fig. S1A). ...
Article
Full-text available
PirB is an inhibitory cell surface receptor particularly prominent on myeloid cells. PirB curtails the phenotypes of activated macrophages during inflammation or tumorigenesis, but its functions in macrophage homeostasis are obscure. To elucidate PirB-related functions in macrophages at steady-state, we generated and compared single-cell RNA-sequencing (scRNAseq) datasets obtained from myeloid cell subsets of wild type (WT) and PirB-deficient knockout (PirB KO) mice. To facilitate this analysis, we developed a novel approach to clustering parameter optimization called “Cluster Similarity Scoring and Distinction Index” (CaSSiDI). We demonstrate that CaSSiDI is an adaptable computational framework that facilitates tandem analysis of two scRNAseq datasets by optimizing clustering parameters. We further show that CaSSiDI offers more advantages than a standard Seurat analysis because it allows direct comparison of two or more independently clustered datasets, thereby alleviating the need for batch-correction while identifying the most similar and different clusters. Using CaSSiDI, we found that PirB is a novel regulator of Cebpb expression that controls the generation of Ly6C lo patrolling monocytes and the expansion properties of peritoneal macrophages. PirB’s effect on Cebpb is tissue-specific since it was not observed in splenic red pulp macrophages (RPMs). However, CaSSiDI revealed a segregation of the WT RPM population into a CD68 lo Irf8 ⁺ “neuronal-primed” subset and an CD68 hi Ftl1 + “iron-loaded” subset. Our results establish the utility of CaSSiDI for single-cell assay analyses and the determination of optimal clustering parameters. Our application of CaSSiDI in this study has revealed previously unknown roles for PirB in myeloid cell populations. In particular, we have discovered homeostatic functions for PirB that are related to Cebpb expression in distinct macrophage subsets.
... An alternative pathway for the development of CD8α + classical DCs in BATF3-/-mice is a cytokine-dependent rescue that is compensated by BATF and BATF2 transcription factors. The positive function of BATFs in the development of DCs is complexation with IRF-4 and IRF-8 (Tussiwand et al. 2012). The non-essentiality of BATF3 in the evolution of CD8-alpha positive DC cells and their cross-presentation has also been mentioned, although these induced cells have short-term survival, which highlights the essential role of BATF3 in increasing cell survival (Seillet et al. 2013). ...
Article
Full-text available
The transcription factor, known as basic leucine zipper ATF-like 3 (BATF3), is a crucial contributor to the development of conventional type 1 dendritic cells (cDC1), which is definitely required for priming CD8 + T cell-mediated immunity against intracellular pathogens and malignancies. In this respect, BATF3-dependent cDC1 can bring about immunological tolerance, an autoimmune response, graft immunity, and defense against infectious agents such as viruses, microbes, parasites, and fungi. Moreover, the important function of cDC1 in stimulating CD8 + T cells creates an excellent opportunity to develop a highly effective target for vaccination against intracellular pathogens and diseases. BATF3 has been clarified to control the development of CD8α⁺ and CD103⁺ DCs. The presence of BATF3-dependent cDC1 in the tumor microenvironment (TME) reinforces immunosurveillance and improves immunotherapy approaches, which can be beneficial for cancer immunotherapy. Additionally, BATF3 acts as a transcriptional inhibitor of Treg development by decreasing the expression of the transcription factor FOXP3. However, when overexpressed in CD8 + T cells, it can enhance their survival and facilitate their transition to a memory state. BATF3 induces Th9 cell differentiation by binding to the IL-9 promoter through a BATF3/IRF4 complex. One of the latest research findings is the oncogenic function of BATF3, which has been approved and illustrated in several biological processes of proliferation and invasion.
... Of interest, BATF3, which functions as a transcription factor, has attracted scientists' attention as a promising therapeutic target for human cancers. BATF3 is highly expressed in conventional Cd11c-positive dendritic cells [18] and it is involved in the homeostatic development of CD8-alpha-positive dendritic cells that trigger the responses of CD8 T-cell against intracellular pathogens [19]. However, BATF3 dysregulation also plays an essential role in the initiation and progression of different types of human cancer. ...
Article
Full-text available
Objective Despite considerable improvement in therapeutic approaches to chronic myeloid leukemia (CML) treatment, this malignancy is considered incurable due to resistance. However, investigating the molecular mechanism of CML may give rise to the development of extremely efficient targeted therapies that improve the prognosis of patients. Basic leucine zipper transcription factor ATF-like3 (BATF3), as transcription factor, is considered a key regulator of cellular activities and its function has been evaluated in tumor development and growth in several cancer types. This study aimed to evaluate the potential of the cellular impact of siRNA-mediated downregulation of BATF3 on CML cancer cells through cell proliferation, induction of apoptosis, and cell cycle distribution. Materials and methods The transfection of BATF3 siRNA to K562 CML cells was performed by electroporation device. To measure cellular viability and apoptosis, MTT assay and Annexin V/PI staining were carried out, respectively. Also, cell cycle assay and flow cytometry instrument were applied to assess cell cycle distribution of K562 cells. For more validation, mRNA expression of correlated genes was relatively evaluated by quantitative real-time polymerase chain reaction (qRT-PCR). Results The data indicated that siRNA-mediated BATF3 inactivating severely promoted the cell apoptosis. Also, the targeted therapy led to high expression of Caspase-3 gene and Bax/Bcl-2 ratio. Silenced BATF3 also induced cell cycle arrest in phase sub-G1 compared to control. Finally, a noticeable decrement was obtained in c-Myc gene expression through suppression of BATF3 in CML cells. Conclusion The findings of this research illustrated the suppression of BATF3 as an effective targeted therapy strategy for CML.
... Old animal-specific TFs were involved in fat cell differentiation, Th17 differentiation and myeloid dendritic cell differentiation pathways, all of which are associated with Batf signaling. [29,30] Signaling associated with Th17 differentiation is known to negatively regulate eosinophil recruitment and IL4 expression, [31][32][33] suggesting that age-associated immunological skewing in the TFs may be responsible for the impaired Th2 response to the regenerative ECM biomaterial. Pathways identified by computational analysis were further validated with biological validation (Figure 1; Figures S4 and S6, Supporting Information). ...
Article
Full-text available
Aging is associated with immunological changes that compromise response to infections and vaccines, exacerbate inflammatory diseases and could potentially mitigate tissue repair. Even so, age‐related changes to the immune response to tissue damage and regenerative medicine therapies remain unknown. Here, we characterized how aging induces changes in immunological signatures that inhibit tissue repair and therapeutic response to a clinical regenerative biological scaffold derived from extracellular matrix. Signatures of inflammation and interleukin (IL)‐17 signaling increased with injury and treatment both locally and regionally in aged animals, and computational analysis uncovered age‐associated senescent‐T cell communication that promotes type 3 immunity in T cells. Local inhibition of type 3 immune activation using IL17‐neutralizing antibodies improved healing and restored therapeutic response to the regenerative biomaterial, promoting muscle repair in older animals. These results provide insights into tissue immune dysregulation that occurs with aging that can be targeted to rejuvenate repair. This article is protected by copyright. All rights reserved
Article
Inflammation and tissue damage associated with pancreatitis can precede or occur concurrently with pancreatic ductal adenocarcinoma (PDAC). We demonstrate that in PDAC coupled with pancreatitis (ptPDAC), antigen-presenting type I conventional dendritic cells (cDC1s) are specifically activated. Immune checkpoint blockade therapy (iCBT) leads to cytotoxic CD8 ⁺ T cell activation and elimination of ptPDAC with restoration of life span even upon PDAC rechallenge. Using PDAC antigen-loaded cDC1s as a vaccine, immunotherapy-resistant PDAC was rendered sensitive to iCBT with elimination of tumors. cDC1 vaccination coupled with iCBT identified specific CDR3 sequences in the tumor-infiltrating CD8 ⁺ T cells with potential therapeutic importance. This study identifies a fundamental difference in the immune microenvironment in PDAC concurrent with, or without, pancreatitis and provides a rationale for combining cDC1 vaccination with iCBT as a potential treatment option.
Article
Full-text available
Clonal expansion of antigen‐specific lymphocytes is the fundamental mechanism enabling potent adaptive immune responses and the generation of immune memory. Accompanied by pronounced epigenetic remodeling, the massive proliferation of individual cells generates a critical mass of effectors for the control of acute infections, as well as a pool of memory cells protecting against future pathogen encounters. Classically associated with the adaptive immune system, recent work has demonstrated that innate immune memory to human cytomegalovirus (CMV) infection is stably maintained as large clonal expansions of natural killer (NK) cells, raising questions on the mechanisms for clonal selection and expansion in the absence of re‐arranged antigen receptors. Here, we discuss clonal NK cell memory in the context of the mechanisms underlying clonal competition of adaptive lymphocytes and propose alternative selection mechanisms that might decide on the clonal success of their innate counterparts. We propose that the integration of external cues with cell‐intrinsic sources of heterogeneity, such as variegated receptor expression, transcriptional states, and somatic variants, compose a bottleneck for clonal selection, contributing to the large size of memory NK cell clones.
Preprint
The basic leucine zipper ATF-like transcription factor (BATF) plays a pivotal role in coordinating various aspects of lymphoid cell biology, yet essential functions in dendritic cells (DCs) have not been reported. Here we demonstrate that BATF deficiency leads to increased interferon (IFN) I production in Toll-like receptor 9 (TLR9)-activated plasmacytoid dendritic cells (pDCs), while BATF overexpression has an inhibitory effect. BATF-deficient mice exhibit elevated IFN I serum levels early in lymphocytic choriomeningitis virus (LCMV) infection. Through ATAC-Seq analysis, BATF emerges as a pioneer transcription factor, regulating approximately one third of the known transcription factors in pDCs. Integrated transcriptomics and ChIP-Seq approaches identified the transcriptional regulator DC-SCRIPT as a direct target of BATF that suppresses IFN I promoter activity by interacting with the interferon regulatory factor 7 (IRF7). Genome-wide association study (GWAS) analyses further implicate BATF in pDC-mediated human diseases. Our findings establish a novel negative feedback axis in IFN I regulation in pDCs during anti-viral immune responses orchestrated by BATF and DC-SCRIPT, with broader implications for pDC and IFN I-mediated autoimmunity.
Article
Background CD4 ⁺ Th1 cells producing IFN-γ are required to eradicate intracellular pathogens, however if uncontrolled these cells can cause immunopathology. The cytokine IL-10 is produced by multiple immune cells including Th1 cells during infection and regulates the immune response to minimise collateral host damage. In this study we aimed to elucidate the transcriptional network of genes controlling the expression of Il10 and proinflammatory cytokines, including Ifng in Th1 cells differentiated from mouse naive CD4 ⁺ T cells. Methods We applied computational analysis of gene regulation derived from temporal profiling of gene expression clusters obtained from bulk RNA sequencing (RNA-seq) of flow cytometry sorted naïve CD4 ⁺ T cells from mouse spleens differentiated in vitro into Th1 effector cells with IL-12 and IL-27 to produce Ifng and Il10, compared to IL-27 alone which express Il10 only , or IL-12 alone which express Ifng and no Il10, or medium control driven-CD4 ⁺ T cells which do not express effector cytokines . Data were integrated with analysis of active genomic regions from these T cells using an assay for transposase-accessible chromatin with sequencing (ATAC)-seq, integrated with literature derived-Chromatin-immunoprecipitation (ChIP)-seq data and the RNA-seq data, to elucidate the transcriptional network of genes controlling expression of Il10 and pro-inflammatory effector genes in Th1 cells. The co-dominant role for the transcription factors, Prdm1 (encoding Blimp-1) and Maf (encoding c-Maf) , in cytokine gene regulation in Th1 cells, was confirmed using T cells obtained from mice with T-cell specific deletion of these transcription factors. Results We show that the transcription factors Blimp-1 and c-Maf each have unique and common effects on cytokine gene regulation and not only co-operate to induce Il10 gene expression in IL-12 plus IL-27 differentiated mouse Th1 cells, but additionally directly negatively regulate key proinflammatory cytokines including Ifng, thus providing mechanisms for reinforcement of regulated Th1 cell responses. Conclusions These data show that Blimp-1 and c-Maf positively and negatively regulate a network of both unique and common anti-inflammatory and pro-inflammatory genes to reinforce a Th1 response in mice that will eradicate pathogens with minimum immunopathology.
Article
Full-text available
The Immunological Genome Project combines immunology and computational biology laboratories in an effort to establish a complete 'road map' of gene-expression and regulatory networks in all immune cells
Article
Full-text available
Interferon regulatory factor 4 (IRF4) and IRF8 regulate B, T, macrophage, and dendritic cell differentiation. They are recruited to cis-regulatory Ets-IRF composite elements by PU.1 or Spi-B. How these IRFs target genes in most T cells is enigmatic given the absence of specific Ets partners. Chromatin immunoprecipitation sequencing in T helper 17 (TH17) cells reveals that IRF4 targets sequences enriched for activating protein 1 (AP-1)–IRF composite elements (AICEs) that are co-bound by BATF, an AP-1 factor required for TH17, B, and dendritic cell differentiation. IRF4 and BATF bind cooperatively to structurally divergent AICEs to promote gene activation and TH17 differentiation. The AICE motif directs assembly of IRF4 or IRF8 with BATF heterodimers and is also used in TH2, B, and dendritic cells. This genomic regulatory element and cognate factors appear to have evolved to integrate diverse immunomodulatory signals.
Article
Full-text available
Interferon regulatory factor 4 (IRF4) is an IRF family transcription factor with critical roles in lymphoid development and in regulating the immune response. IRF4 binds DNA weakly owing to a carboxy-terminal auto-inhibitory domain, but cooperative binding with factors such as PU.1 or SPIB in B cells increases binding affinity, allowing IRF4 to regulate genes containing ETS-IRF composite elements (EICEs; 5'-GGAAnnGAAA-3'). Here we show that in mouse CD4(+) T cells, where PU.1/SPIB expression is low, and in B cells, where PU.1 is well expressed, IRF4 unexpectedly can cooperate with activator protein-1 (AP1) complexes to bind to AP1-IRF4 composite (5'-TGAnTCA/GAAA-3') motifs that we denote as AP1-IRF composite elements (AICEs). Moreover, BATF-JUN family protein complexes cooperate with IRF4 in binding to AICEs in pre-activated CD4(+) T cells stimulated with IL-21 and in T(H)17 differentiated cells. Importantly, BATF binding was diminished in Irf4(-/-) T cells and IRF4 binding was diminished in Batf(-/-) T cells, consistent with functional cooperation between these factors. Moreover, we show that AP1 and IRF complexes cooperatively promote transcription of the Il10 gene, which is expressed in T(H)17 cells and potently regulated by IL-21. These findings reveal that IRF4 can signal via complexes containing ETS or AP1 motifs depending on the cellular context, thus indicating new approaches for modulating IRF4-dependent transcription.
Article
Full-text available
Distinguishing dendritic cells (DCs) from other cells of the mononuclear phagocyte system is complicated by the shared expression of cell surface markers such as CD11c. In this study, we identified Zbtb46 (BTBD4) as a transcription factor selectively expressed by classical DCs (cDCs) and their committed progenitors but not by plasmacytoid DCs (pDCs), monocytes, macrophages, or other lymphoid or myeloid lineages. Using homologous recombination, we replaced the first coding exon of Zbtb46 with GFP to inactivate the locus while allowing detection of Zbtb46 expression. GFP expression in Zbtb46(gfp/+) mice recapitulated the cDC-specific expression of the native locus, being restricted to cDC precursors (pre-cDCs) and lymphoid organ- and tissue-resident cDCs. GFP(+) pre-cDCs had restricted developmental potential, generating cDCs but not pDCs, monocytes, or macrophages. Outside the immune system, Zbtb46 was expressed in committed erythroid progenitors and endothelial cell populations. Zbtb46 overexpression in bone marrow progenitor cells inhibited granulocyte potential and promoted cDC development, and although cDCs developed in Zbtb46(gfp/gfp) (Zbtb46 deficient) mice, they maintained expression of granulocyte colony-stimulating factor and leukemia inhibitory factor receptors, which are normally down-regulated in cDCs. Thus, Zbtb46 may help enforce cDC identity by restricting responsiveness to non-DC growth factors and may serve as a useful marker to identify rare cDC progenitors and distinguish between cDCs and other mononuclear phagocyte lineages.
Article
Full-text available
Cancer immunoediting, the process by which the immune system controls tumour outgrowth and shapes tumour immunogenicity, is comprised of three phases: elimination, equilibrium and escape. Although many immune components that participate in this process are known, its underlying mechanisms remain poorly defined. A central tenet of cancer immunoediting is that T-cell recognition of tumour antigens drives the immunological destruction or sculpting of a developing cancer. However, our current understanding of tumour antigens comes largely from analyses of cancers that develop in immunocompetent hosts and thus may have already been edited. Little is known about the antigens expressed in nascent tumour cells, whether they are sufficient to induce protective antitumour immune responses or whether their expression is modulated by the immune system. Here, using massively parallel sequencing, we characterize expressed mutations in highly immunogenic methylcholanthrene-induced sarcomas derived from immunodeficient Rag2(-/-) mice that phenotypically resemble nascent primary tumour cells. Using class I prediction algorithms, we identify mutant spectrin-β2 as a potential rejection antigen of the d42m1 sarcoma and validate this prediction by conventional antigen expression cloning and detection. We also demonstrate that cancer immunoediting of d42m1 occurs via a T-cell-dependent immunoselection process that promotes outgrowth of pre-existing tumour cell clones lacking highly antigenic mutant spectrin-β2 and other potential strong antigens. These results demonstrate that the strong immunogenicity of an unedited tumour can be ascribed to expression of highly antigenic mutant proteins and show that outgrowth of tumour cells that lack these strong antigens via a T-cell-dependent immunoselection process represents one mechanism of cancer immunoediting.
Article
Full-text available
Dendritic cells (DCs) subsets differ in precursor cell of origin, functional properties, requirements for growth factors, and dependence on transcription factors. Lymphoid-tissue resident CD8α(+) conventional DCs (cDCs) and CD11b(low/-)CD103(+) non-lymphoid DCs are developmentally related, each being dependent on FMS-like tyrosine kinase 3 ligand (Flt3L), and requiring the transcription factors Batf3, Irf8, and Id2 for development. It was recently suggested that granulocyte/macrophage colony stimulating factor (GM-CSF) was required for the development of dermal CD11b(low/-)Langerin(+)CD103(+) DCs, and that this dermal DC subset was required for priming autoreactive T cells in experimental autoimmune encephalitis (EAE). Here, we compared development of peripheral tissue DCs and susceptibility to EAE in GM-CSF receptor deficient (Csf2rb(-/-)) and Batf3(-/-) mice. We find that Batf3-dependent dermal CD11b(low/-)Langerin(+) DCs do develop in Csf2rb(-/-) mice, but that they express reduced, but not absent, levels of CD103. Further, Batf3(-/-) mice lacking all peripheral CD11b(low/-) DCs show robust Th cell priming after subcutaneous immunization and are susceptible to EAE. Our results suggest that defective T effector priming and resistance to EAE exhibited by Csf2rb(-/-) mice does not result from the absence of dermal CD11b(low/-)Langerin(+)CD103(+) DCs.
Article
Interferon regulatory factor 4 (IRF4) and IRF8 regulate B, T, macrophage, and dendritic cell differentiation. They are recruited to cis-regulatory Ets-IRF composite elements by PU.1 or Spi-B. How these IRFs target genes in most T cells is enigmatic given the absence of specific Ets partners. Chromatin immunoprecipitation sequencing in T helper 17 (T(H)17) cells reveals that IRF4 targets sequences enriched for activating protein 1 (AP-1)-IRF composite elements (AICEs) that are co-bound by BATF, an AP-1 factor required for T(H)17, B, and dendritic cell differentiation. IRF4 and BATF bind cooperatively to structurally divergent AICEs to promote gene activation and T(H)17 differentiation. The AICE motif directs assembly of IRF4 or IRF8 with BATF heterodimers and is also used in T(H)2, B, and dendritic cells. This genomic regulatory element and cognate factors appear to have evolved to integrate diverse immunomodulatory signals.
Article
We generated mice harboring germline mutations in which the enhancer element located 9 kb 3′ of the immunoglobulin κ light chain gene (3′Eκ) was replaced either by a single loxP site (3′EκΔ) or by a neomycin resistance gene (3′EκN). Mice homozygous for the 3′EκΔ mutation had substantially reduced numbers of κ-expressing B cells and increased numbers of λ-expressing B cells accompanied by decreased κ versus λ gene rearrangement. In these mutant mice, κ expression was reduced in resting B cells, but was normal in activated B cells. The homozygous 3′EκN mutation resulted in a similar but more pronounced phenotype. Both mutations acted in cis. These studies show that the 3′Eκ is critical for establishing the normal κ/λ ratio, but is not absolutely essential for κ gene rearrangement or, surprisingly, for normal κ expression in activated B cells. These studies also imply the existence of additional regulatory elements that have overlapping function with the 3′Eκ element.
Article
The nuclear factor of activated T cells (NFAT) and the AP-1 heterodimer, Fos-Jun, cooperatively bind a composite DNA site and synergistically activate the expression of many immune-response genes. A 2.7-A-resolution crystal structure of the DNA-binding domains of NFAT, Fos and Jun, in a quaternary complex with a DNA fragment containing the distal antigen-receptor response element from the interleukin-2 gene promoter, shows an extended interface between NFAT and AP-1, facilitated by the bending of Fos and DNA. The tight association of the three proteins on DNA creates a continuous groove for the recognition of 15 base pairs.
Article
CD8α(+) dendritic cells (DCs) are important in vivo for cross-presentation of antigens derived from intracellular pathogens and tumors. Additionally, secretion of interleukin-12 (IL-12) by CD8α(+) DCs suggests a role for these cells in response to Toxoplasma gondii antigens, although it remains unclear whether these cells are required for protection against T. gondii infection. Toward this goal, we examined T. gondii infection of Batf3(-/-) mice, which selectively lack only lymphoid-resident CD8α(+) DCs and related peripheral CD103(+) DCs. Batf3(-/-) mice were extremely susceptible to T. gondii infection, with decreased production of IL-12 and interferon-γ. IL-12 administration restored resistance in Batf3(-/-) mice, and mice in which IL-12 production was ablated only from CD8α(+) DCs failed to control infection. These results reveal that the function of CD8α(+) DCs extends beyond a role in cross-presentation and includes a critical role for activation of innate immunity through IL-12 production during T. gondii infection.