ArticlePDF Available

A Costimulatory Function for T Cell CD40

Authors:
  • Oklahoma Medical Research Foundation and Progentec Diagnostics

Abstract and Figures

CD40 plays a significant role in the pathogenesis of inflammation and autoimmunity. B cell CD40 directly activates cells, which can result in autoantibody production. T cells can also express CD40, with an increased frequency and amount of expression seen in CD4(+) T lymphocytes of autoimmune mice, including T cells from mice with collagen-induced arthritis. However, the mechanisms of T cell CD40 function have not been clearly defined. To test the hypothesis that CD40 can serve as a costimulatory molecule on T lymphocytes, CD40(+) T cells from collagen-induced arthritis mice were examined in parallel with mouse and human T cell lines transfected with CD40. CD40 served as effectively as CD28 in costimulating TCR-mediated activation, including induction of kinase and transcription factor activities and production of cytokines. An additional enhancement was seen when both CD40 and CD28 signals were combined with AgR stimulation. These findings reveal potent biologic functions for T cell CD40 and suggest an additional means for amplification of autoimmune responses.
Content may be subject to copyright.
A Costimulatory Function for T Cell CD40
1
Melissa E. Munroe* and Gail A. Bishop
2
*
†‡
CD40 plays a significant role in the pathogenesis of inflammation and autoimmunity. B cell CD40 directly activates cells, which
can result in autoantibody production. T cells can also express CD40, with an increased frequency and amount of expression seen
in CD4
T lymphocytes of autoimmune mice, including T cells from mice with collagen-induced arthritis. However, the mech
-
anisms of T cell CD40 function have not been clearly defined. To test the hypothesis that CD40 can serve as a costimulatory
molecule on T lymphocytes, CD40
T cells from collagen-induced arthritis mice were examined in parallel with mouse and human
T cell lines transfected with CD40. CD40 served as effectively as CD28 in costimulating TCR-mediated activation, including
induction of kinase and transcription factor activities and production of cytokines. An additional enhancement was seen when both
CD40 and CD28 signals were combined with AgR stimulation. These findings reveal potent biologic functions for T cell CD40 and
suggest an additional means for amplification of autoimmune responses. The Journal of Immunology, 2007, 178: 671– 682.
T
he 50-kDa membrane receptor of the TNFR superfamily,
CD40, is expressed by APC, including dendritic cells,
macrophages, and B lymphocytes (1, 2). The ligand for
CD40, CD154, is expressed on activated T cells and allows for
interactions with APC during the cognitive phase of the immune
response, as well as directing effector T cell-dependent B cell ac-
tivation (3). Such interactions have been directly implicated in
autoimmunity. Blocking CD40-CD154 interactions has been
shown to either prevent or alleviate such diseases as insulin de-
pendent diabetes mellitus (4), arthritis (5), and systemic lupus er-
ythematosus (6), the latter two benefiting from abrogated T cell-
dependent autoantibody production (7, 8).
The role of CD40 as a direct signal receptor has now been ex-
panded to T cells. Shortly after CD154 was cloned, it was dem-
onstrated that CD154 can augment mitogen and TCR-mediated
proliferation of CD4
and CD8
T cells (9), although the lack of
CD154-specific Abs at that time precluded further investigation.
Although a specific biologic role for CD40 on CD8
T cells re
-
mains undefined (10 –13), it has been shown that autoimmune-
prone strains of mice have increased numbers of CD40
CD4
T
cells compared with normal strains (14). The most extensively
studied of these is the NOD mouse (15–17).
To date, the physiologic role(s) of CD40 on T cells has not been
characterized, nor have the mechanisms by which CD40 affects T
cell function been defined. We have previously studied CD40 as an
important signaling molecule on B lymphocytes, delivering signals
alone and synergistically with the BCR (18), leading to NF-
B and
JNK pathway activation and subsequent proliferation, secretion of
cytokines, Ig production and isotype switching (reviewed in Refs.
19 and 20). Like B cells, T cells have been shown to use TNFR
family members as costimulatory molecules for Ag receptor stim-
ulation (reviewed in Refs. 21 and 22). We hypothesized that co-
stimulation is a plausible role for CD40 on T cells. In the studies
presented here, we determined that T cell CD40 augmented CD3
and CD3 plus CD28-mediated cytokine production in T cells from
mice which have developed collagen-induced arthritis (CIA),
3
as
well as in T cell lines stably transfected with CD40. Although
CD40 signals alone did not activate NFAT or IL-2 secretion, CD40
ligation markedly augmented CD3 and CD3 plus CD28 responses.
As in B cells, T cell CD40 was able to efficiently bind the adaptor
proteins TNFR-associated factors (TRAF), activate NF-
B and
AP-1 pathways, and stimulate TNF-
secretion. Taken together,
these findings reveal that CD40 can act as a powerful signaling
receptor on T as well as B lymphocytes, a function that may have
important implications for T-B interactions in autoimmune
diseases.
Materials and Methods
Cells
The mouse T cell line 2B4.11 (23) and human T cell line Jurkat (24) have
been described previously. Cell lines and their stable transfectants express-
ing hCD40 were maintained in RPMI 1640 containing 10% FCS (Hy-
Clone), 10
M 2-ME, and antibiotics. These subclones are referred to as
2B4.hCD40 and J.hCD40. Hi5 insect cells expressing hCD154 have been
described and characterized previously (25, 26). These cells grow at 26°C
and rapidly die to form membrane fragments at 37°C and therefore do not
overgrow cell cultures.
Stable transfections
Cell lines were stably transfected with a previously reported hCD40 ex-
pression plasmid (27) as described previously (28). G418-resistant clones
were analyzed for expression of hCD40 using a FACScan flow cytometer
(BD Biosciences) and mean channel fluorescence (MCF) determined using
FlowJo software.
Reagents
Recombinant mouse TNF-
, IL-2, and IFN-
were purchased from Pep-
roTech. Streptavidin-HRP was purchased from Jackson ImmunoResearch
Laboratories. ELISA TMB peroxidase substrate was purchased from KPL.
Tosylactivated Dynabeads for Ab conjugation (per manufacturer’s instruc-
tions) were purchased from Dynal Biotech. PMA and ionomycin were
purchased from Sigma-Aldrich.
*Department of Microbiology and
Department of Internal Medicine, University of
Iowa, Iowa City, IA 52242; and
Veterans Affairs Medical Center, Iowa City, IA
52242
Received for publication June 21, 2006. Accepted for publication October 17, 2006.
The costs of publication of this article were defrayed in part by the payment of page
charges. This article must therefore be hereby marked advertisement in accordance
with 18 U.S.C. Section 1734 solely to indicate this fact.
1
This work was supported by grants from the National Institutes of Health and the
Veterans’ Administration (to G.A.B.) and postdoctoral fellowship support provided
by the American Heart Association and the American Cancer Society (to M.E.M.).
2
Address correspondence and reprint requests to Dr. Gail A. Bishop, 2193B MERF,
Department of Microbiology, University of Iowa, Iowa City, IA 52242. E-mail ad-
dress: gail-bishop@uiowa.edu
3
Abbreviations used in this paper: CIA, collagen-induced arthritis; CII, type II
chicken collagen; MCF, mean channel fluorescence; TRAF, TNFR-associated factor.
Copyright © 2007 by The American Association of Immunologists, Inc. 0022-1767/07/$2.00
The Journal of Immunology
www.jimmunol.org
Antibodies
The 1C10 (anti-mCD40, rat IgG2a), 72-2 (rat IgG2a isotype control), and
G28-5 (anti-hCD40, mouse IgG1) hybridomas were purchased from the
American Type Culture Collection. MOPC-31c (mouse IgG1 isotype con-
trol) was from Sigma-Aldrich. Polyclonal rabbit anti-TRAF2 Ab was from
MBL. Polyclonal mouse anti-yy1 and polyclonal rabbit anti-TRAF3, anti-
TRAF1, anti-TRAF6, and anti-hCD40 Abs were from Santa Cruz Biotech-
nology. Polyclonal rabbit anti-I
B
, anti-phosphorylated I
B
, anti-
NF
B2 p100/p52, anti-JNK, and anti-phosphorylated JNK Abs were from
Cell Signaling Technology. Mouse anti-actin Ab (C4) was from Chemicon
International. Peroxidase-labeled goat anti-rabbit and goat anti-mouse IgG
Abs were from Jackson ImmunoResearch Laboratories. Anti-mouse CD3
(145-2C11; Armenian hamster IgG1), anti-human CD3 (OKT3; mouse
IgG2a), anti-mouse CD28 (37.51; hamster IgG), anti-human CD28
(CD28.2; mouse IgG1), and relevant isotype control Abs were purchased
from eBioscience. PE-labeled anti-mouse CD3 (145-2C11; Armenian ham-
ster IgG1), FITC labeled anti-mouse CD40 (HM40-3; Armenian hamster
IgM), anti-human CD40 (5C3; mouse IgG3), anti-human CD154 (TRAP1,
mouse IgG1), anti-mouse CD80 (16-10A1; Armenian hamster IgG2), anti-
mouse CD86 (GL1; rat IgG2a), anti-mouse CD95 (Jo2; Armenian hamster
IgG2), and relevant isotype control Abs were purchased from BD Pharm-
ingen. FITC labeled anti-mouse CD25 (PC61.5; rat IgG1), anti-mouse
CD54 (YN1/1.7.4; rat IgG2b), and anti-mouse CD11
(M17/4; rat IgG2a)
Abs were purchased from eBioscience. Biotin-labeled anti-mouse CD154
(MR1; Armenian hamster IgG) and relevant isotype control Abs were pur-
chased from eBioscience. Alexa Fluor 488-labeled streptavidin was pur-
chased from Molecular Probes/Invitrogen Life Technologies. Anti-mouse
IL-2 and IFN-
(coating and biotinylated) ELISA Abs were purchased
from Caltag Laboratories. Anti-human IL-2 and anti-mouse TNF-
(coat-
ing and biotinylated) ELISA Abs were purchased from eBioscience.
Mice/CIA induction
Female C57BL/6 mice were purchased at 5– 8 wk of age from the National
Cancer Institute. Mice were housed in a specific pathogen-free barrier fa-
cility with restricted access, and all procedures were performed as ap-
proved by the University of Iowa Animal Care and Use Committee. CIA
was induced based on the methods of Campbell et al. (29). Briefly, mice
were either left naive, immunized in the tail s.c. with 100
g of type II
chicken collagen (CII; Sigma-Aldrich) dissolved in 10 mM acetic acid and
emulsified in IFA (Sigma-Aldrich) containing 5 mg/ml H37 RA heat-killed
mycobacteria (CFA; Difco Laboratories), or immunized with 10 mM acetic
acid emulsified in CFA. Mice were monitored for limb erythema and swell-
ing and paws measured (each paw recorded individually; four measure-
ments per mouse) with calipers two to three times per week (4026F; Mi-
tutoyo) (29, 30). All mouse studies were reviewed and approved by the
University of Iowa Animal Care and Use Committee.
T cell isolation
T cells were isolated from mouse spleens 70 days postimmunization.
Briefly, spleens from euthanized mice were teased apart with forceps,
erythrocytes lysed in ACK buffer, and remaining cells placed over a T cell
enrichment column per manufacturer’s protocol (R&D Biosystems). T
cells were enriched to 90% purity as determined by flow cytometry (see
Table I). Primary mouse T cells were cultured in Click’s medium contain-
ing 1% nutridoma-SP (Roche), 10
M 2-ME, and antibiotics, or stained for
CD3 (PE) and CD40 (FITC) and analyzed by flow cytometry.
NF-
B/NFAT/AP-1 dual luciferase reporter assays
2B4.hCD40 or J.hCD40 cells (1.5 10
7
) were transiently transfected with
20
gof4 NF-
B, 40
gof4 NFAT, or 40
gof7 AP-1 luciferase
reporter plasmid (31), and 1
gofRenilla luciferase vector (pRL-null;
Promega) by electroporation. Cells were rested on ice for 15 min, then
stimulated (2 10
6
cells/ml) for 6 h (NF-
B) or 24 h (NFAT/AP-1) with
10
g/ml anti-hCD40 or isotype controls, and/or 5 10
5
beads/ml anti-
CD3 or anti-CD3
CD28-coated Dynabeads. After stimulation, cells were
pelleted, lysed, and assayed for relative luciferase activity (NF-
B, NFAT,
or AP-1:Renilla) per manufacturer’s protocol (Promega) using a Turner
Designs 20/20 luminometer, with settings of a 2-s delay followed by a
10-s read.
I
B
/JNK assays
2B4.hCD40 or J.hCD40 cells (2 10
6
) were stimulated for indicated times
with culture medium, 10
g of anti-hCD40 Ab (or respective isotype con-
trols) and/or 5 10
5
beads/ml anti-CD3 or anti-CD3 plus CD28-coated
Dynabeads) to induce phosphorylation and degradation of the proteins
blotted. The cells were pelleted by centrifugation, lysed and analyzed by
SDS PAGE and Western blotting. Peroxidase-labeled Abs were visualized
on Western blots using a chemiluminescent detection reagent (Pierce).
NF-
B2 activation
2B4.hCD40 or J.hCD40 cells (2 10
6
) were stimulated for indicated times
with culture medium, 10
g of anti-hCD40 Ab (or respective isotype con-
trols) and/or 5 10
5
beads/ml anti-CD3 or anti-CD3 plus CD28-coated
Dynabeads) to induce RelB activation, and processing of p100 to p52. The
cells were pelleted by centrifugation and cytoplasmic and nuclear fractions
isolated as described previously (32). Samples were analyzed by SDS-
PAGE and Western blotting. Peroxidase-labeled Abs were visualized on
Western blots using a chemiluminescent detection reagent.
FIGURE 1. CD40 expressed on T cells from mice
with CIA. A, C57BL/6 mice were immunized with CFA
only, CII plus CFA, or remained naive, as described in
Materials and Methods. Paws were measured two to
three times per week from days 21–70 postimmuniza-
tion. Mice receiving CII plus CFA had significant (p
0.001) paw swelling compared with CFA or naive con-
trols. Data represent two experiments (4 mice/group/
experiment; 8 mice/32 paws/data point total). B,
Spleens from mice in A were pooled (2 spleens/pool, 4
pools/group), and T cells were isolated as described in
Materials and Methods. Cells were stained with anti-
mouse CD40 (FITC) and anti-mouse CD3 (PE) or iso-
type control Abs and analyzed by flow cytometry.
Quadrants were drawn based on staining with isotype
control Abs.
672 SIGNALING BY CD40 IN T CELLS
Cytokine ELISA
Primary mouse T cells or 2B4.hCD40 or J.hCD40 cells (4 10
5
) were
stimulated at optimal, empirically derived time points with culture me-
dium, 1
g/ml anti-hCD40 Ab and/or anti-CD28, or plate-bound anti-CD3
(or respective isotype controls). Cytokine concentrations in culture super-
natants were determined by ELISA, using cytokine-specific coating Abs
and biotinylated detection Abs. Streptavidin-HRP binding to biotinylated
detection Abs was visualized with TMB substrate and the reaction was
stopped with 0.18 M H
2
SO
4
. Plates were read at 450 nm by a Spectra
-
Max
250
Reader (Molecular Devices). Data were analyzed with SoftMax
Pro software (Molecular Devices); unknowns were compared with a stan-
dard curve containing at least five to seven dilution points of the relevant
recombinant cytokine on each assay plate. In all cases, the coefficient of
determination for the standard curve (r
2
) was 0.98. ELISA unknowns
were diluted to fall within the standard values.
TRAF recruitment to receptors in detergent-insoluble
microdomains (Rafts) and immunoprecipitation.
2B4.hCD40 or J.hCD40 (1 10
7
) cells were stimulated with 10
g of anti-hCD40 Ab (or isotype control Abs) or Hi5 cells ex-
pressing hCD154 (or Hi5 cells expressing WT baculovirus; 1:4
Hi5 cells:lymphocytes) for 15 min at 37°C to induce recruitment of
TRAFs to membrane rafts and allow formation of CD40 signaling
complexes, as described previously (33). Detergent (1% Brij 58)-
soluble and insoluble fractions were separated as described previ-
ously (34). Samples of soluble and insoluble lysates were reserved
for SDS-PAGE separation and analysis by Western blotting. The
remainder of the lysates were immunoprecipitated with protein
G-Sepharose beads (Amersham Biosciences) prewarmed with anti-
hCD40 Ab for3hat4°C. The immunoprecipitation complexes
were washed four times with lysis buffer before separation by
SDS-PAGE and analysis by Western blot.
Up-regulation of cell surface proteins
In experiments evaluating activation-induced up-regulation of surface pro-
teins, 2B4.11 or 2B4.hCD40 cells were incubated with the indicated stimuli
in 96-well plates (1–2 10
5
cells/well). After 48–72 h, the cells were
washed, then incubated for 20 min on ice in PBS-0.5% FCS-0.02% sodium
azide containing 2.5 mM EDTA. EDTA treatment helped to dissociate cell
aggregates formed upon CD40 stimulation. Following the EDTA incuba-
tions, cells were washed and stained (in the absence of EDTA) with Abs for
analysis by flow cytometry.
Statistical analyses
Analyses were performed with GraphPad Instat software. A two-tailed
paired Student’s t test was used to determine significance between groups
in CIA experiments, for cytokine ELISA, surface molecule up-regulation
experiments, and luciferase reporter assays.
Results
CD40 expression on T cells from mice with CIA
CIA is a frequently used mouse model of inflammatory rheumatoid
arthritis that is both Ab and T cell dependent (35, 36). It has been
previously demonstrated that mouse strains prone to autoimmune
diabetes have an increased number of T cells expressing CD40 that
correlates with development of pathology (14). We examined
FIGURE 2. CD40 enhances CD3 and CD3 plus CD28-mediated cyto-
kine production in T cells from mice with CIA. Splenic T cells from
C57BL/6 mice immunized with CII plus CFA, CFA only, or remaining
naive were isolated as described in Materials and Methods. Cells were
stimulated with plate-bound anti-CD3 anti-CD28 and/or anti-mouse
CD40 Abs (compared with medium (Med)/isotype control (IC)). Culture
supernatants were collected and assayed for IL-2 (A, 48 h), IFN-
(B,72h),
or TNF-
(C, 24 h) by ELISA. Data represent the mean SEM of du-
plicate samples of three independent experiments. There was no difference
between medium and isotype control samples.
Table I. Increased number of CD40
T cells in CIA mice
a
CII/CFA CFA Only Naive
Splenocytes (pre-T cell isolation)
Percentage of CD3
T cells
33.19 0.63 33.52 0.28 34.44 0.58
No. of CD3
T cells (10
6
)
5.67 0.17 5.11 0.17 4.81 0.35
Percentage of CD3/CD40
T cells
9.31 0.52
b
4.30 0.37 4.02 0.46
No. of CD3/CD40
T cells (10
5
)
15.90 1.54
c
6.55 0.60 6.06 0.85
Negatively selected T cells
Percentage of CD3
T cells
88.67 1.56 87.41 3.28 89.01 1.67
No. of CD3
T cells (10
6
)
3.85 1.75 3.24 0.30 3.68 0.34
Percentage of CD3/CD40
T cells
7.53 0.70
d
2.96 0.26 2.80 0.56
No. of CD3/CD40
T cells (10
5
)
3.31 0.48
e
1.11 0.17 1.40 0.11
a
T cells were isolated as described in Materials and Methods from spleens of C57BL/6 mice receiving CII/CFA, CFA only,
or remaining naive (four mice per group for two experiments; two spleens per pool, four pools total) 70 days postimmunization.
Cells were stained with anti-mouse CD40 (FITC) and anti-mouse CD3
(PE) or isotype control mAbs and analyzed by flow
cytometry. Spleens from CII/CFA mice had significantly more CD3
CD40
T cells than either CFA or naïve controls.
b
p 0.01 (CII/CFA compared with CFA only or naive).
c
p 0.001 (CII/CFA compared with CFA only or naive).
d
p 0.01 (CII/CFA compared with CFA only or naive).
e
p 0.05 (CII/CFA compared with CFA only or naive).
673The Journal of Immunology
whether this was also true during the inflammatory process of CIA
development. C57BL/6 mice injected with CII/CFA developed
significant paw swelling ( p 0.001) compared with mice given
CFA only or naive controls (Fig. 1A). Splenic T cells from these
mice were isolated and evaluated for dual expression of CD3 and
CD40. A representative FACS plot of isolated T cells is presented
in Fig. 1B; quantitation is presented in Table I. Small numbers of
cells in the FACS samples that were CD3
were also CD40
and
therefore unlikely to be APC (Fig. 1B and Table I, line 5). Data in
Table I demonstrate that there are more than twice as many CD40
T cells in the spleens of mice that received CII/CFA compared
with mice that received CFA only or naive controls ( p 0.01 for
percentage, line 3, or absolute number, line 4). This is true whether
evaluating the percentage or absolute numbers of CD3
CD40
cells as a part of the whole spleen (Table I, lines 3 and 4) or after
enrichment of T lymphocytes (Table I, lines 7 and 8). This phe-
nomenon was not due to an alteration in the splenic T cell popu-
lation as a whole, or post-T cell isolation after CII/CFA immuni-
zation. There was no significant difference between immunization
groups in the percentage (first line) or absolute numbers (second
line) of CD3
T cells, whether evaluating mixed splenocyte, or
isolated T cell populations (Table I).
T cell CD40 as a costimulatory molecule
The findings discussed above raise the possibility that CD40 ex-
pressed by T cells may play a role in T cell activation. Experiments
presented in Fig. 2 explored whether CD40 engagement could aug-
ment CD3 or CD3 plus CD28-mediated cytokine production by
splenic T cells isolated from mice immunized with CII/CFA. Be-
cause minimal cytokine production was detected from cells stim-
ulated with medium alone, isotype control Abs, or anti-CD3 Ab, it
is unlikely that the small amounts of residual non-T cells found
after T cell enrichment (Fig. 1 and Table I) are APC. As expected,
anti-CD3 Ab induced a modest amount of IL-2 (Fig. 2A), and
anti-CD28 Ab significantly enhanced CD3-mediated IL-2 produc-
tion in all three experimental groups (CD3 vs CD3 plus CD28:
naive, p 0.02; CFA only, p 0.01; CII/CFA, p 0.0001).
CD40 stimulation alone did not induce any appreciable IL-2 pro-
duction. However, in T cells from mice immunized with CII/CFA,
CD40 significantly enhanced the level of CD3 ( p 0.002)- and
CD3 plus CD28 ( p 0.009)-mediated IL-2 production. This en-
hancement did not occur in the largely CD40-negative T cells from
naive or CFA-treated mice because there was no significant dif-
ference in IL-2 produced between T cells treated with agonists for
CD3 vs CD3 plus CD40 or CD3 plus CD28 vs CD3 plus CD28
plus CD40.
In addition to IL-2 production, we also evaluated the ability of
CD40 to contribute to the production of the proinflammatory cy-
tokines IFN-
(Fig. 2B) and TNF-
(Fig. 2C) by T cells from CIA
mice compared with controls. As with IL-2 secretion, CD28 en-
hanced CD3-mediated IFN-
(Fig. 2B, p 0.001 for all groups)
and TNF-
(Fig. 2C, p 0.001 for all groups) production in T
cells from all three mouse groups. Unlike IL-2, anti-CD40 alone
induced a significant amount of both IFN-
(Fig. 2B, p 0.001)
and TNF-
(Fig. 2C, p 0.0001), but only in T cell cultures from
mice immunized with CII/CFA and not in cultures from control
mice. CD40 significantly enhanced CD3 and CD3 plus CD28-me-
diated IFN-
(Fig. 2B, p 0.02 for both stimuli) and TNF-
( p
0.01 for both stimuli) production, but only in mice immunized with
CII/CFA. These data indicate that CD40 can act as a TCR co-
stimulator and that it can cooperate in a nonredundant manner with
CD28 to further enhance T cell cytokine production.
The above ex vivo experiments contained a mixed population of
CD40-expressing and nonexpressing T cells (Fig. 2), with a rela-
tively small percentage of T cells expressing CD40 (Fig. 1 and
Table I). This small percentage limited detailed molecular charac-
terization of CD40 function on CD40-expressing T cells, although
the data presented in Fig. 2 indicate that this population has sig-
nificant biologic activity distinct from that of CD4
T cells that do
not express CD40. We thus wanted to complement these experi-
ments with stimulation of homogeneous populations of CD40
T
cells, as well as explore CD40 signaling pathways. Because of
limiting numbers of CD40
T cells in CIA mice that would require
potentially function-altering positive selection for isolation,
CD40
T cell lines were a desirable alternative. We thus stably
transfected mouse 2B4.11 (2B4.hCD40) and human Jurkat
(J.hCD40) T cell lines with hCD40 (Fig. 3, A and B; MCF for
2B4.hCD40 704.66 vs 147.48 for 2B4.11; MCF for J.hCD40
245.72 vs 199.11 for Jurkat) and evaluated the ability of CD40 to
activate IL-2 production (Fig. 4, A and B). Interestingly, even with
PMA/ionomycin stimulation, there was no detectable CD154 ex-
pression by either 2B4.hCD40 or J.hCD40 cells (Fig. 3, C and D),
although we see CD154 expression by Hi5 insect cells infected
with a baculovirus encoded to express CD154 (data not shown).
Because these clones do not express CD154, autocrine CD40 stim-
ulation does not contribute to subsequent findings.
We first evaluated the ability of the transfected 2B4.hCD40 and
J.hCD40 to secrete cytokines in response to CD3 CD28 and/or
CD40 stimulation (Fig. 4). To closely mimic the design of our
mouse ex vivo experiments, in which biologic responses to CD40
FIGURE 3. Expression of hCD40 and CD154 on 2B4.11 and Jurkat cell
lines. A and B, 2B4.11 (A) or Jurkat (B) cells were transfected with DNA
encoding for hCD40. Cells were stained with anti-human CD40 and ana-
lyzed by flow cytometry. Gray profiles represent anti-hCD40 mAb staining
of untransfected cells; black profiles represent staining of transfected cells.
There was no significant staining of cells with an isotype control Ab. C and
D, 2B4.hCD40 (C) or J.hCD40 (D) cells remained untreated or were
treated for 6 h with 1
g/ml PMA plus 200
M ionomycin, then stained
with anti-CD154 and analyzed by flow cytometry. Dashed profiles repre-
sent staining of untreated cells with isotype control mAb; gray profiles
represent anti-CD154 staining of untreated cells; black profiles represent
anti-CD154 staining of PMA-ionomycin-treated cells. There was no sig-
nificant staining of PMA-ionomycin-treated cells with an isotype control
Ab. Similar results were seen at 24 and 48 h (data not shown).
674 SIGNALING BY CD40 IN T CELLS
were made by small percentages of CD40
T cells, CD40 trans
-
fected T cells were cocultured with their untransfected parent cell
lines at various ratios ranging from 6 to 100% transfected cells.
CD40 stimulation alone of cells stably expressing the transfected
receptor did not induce IL-2 production (Fig. 4, A and B). While
CD28 ligation appropriately enhanced CD3-mediated IL-2 produc-
tion in all T cell clones, CD40 stimulation enhanced IL-2 produc-
tion via CD3 or CD3 plus CD28 only if cells stably expressing
CD40 were present in the culture well, similar to responses of
primary T cell cultures from CIA mice (Fig. 2B). CD40 was able
to act in costimulatory fashion with CD3 with only 6% of CD40
cells in the coculture ( p 0.01 when 6% 2B4.hCD40 present, p
0.0001 at 100% 2B4.hCD40; p 0.001 when 6% J.hCD40
present, p 0.0001 at 100% J.hCD40), just as was seen in primary
T cells from CIA mice (Fig. 2B). Enhancement of CD3 plus CD28-
mediated IL-2 production by CD40 increased with greater num-
bers of transfected cells in the well ( p 0.02 at 100%
2B4.hCD40; p 0.005 when 6% J.hCD40 present, p 0.0001 at
100% J.hCD40). Of particular interest is that CD40 could ulti-
mately enhance CD3-mediated IL-2 production to a degree similar
to that of CD28 in 2B4.hCD40 cells (Fig. 4A) and significantly
better in J.hCD40 cells (Fig. 4B, p 0.0001 at 100% J.hCD40).
Thus, the response of both 2B4.hCD40 and J.hCD40, even when
only 6% of the T cells in each culture were CD40
cells, closely
paralleled the responses seen in freshly isolated T cells, validating
these cell lines as useful experimental models for the study of
CD40 function in T cells.
In addition to IL-2 production, we evaluated the ability of CD40
to contribute to the production of proinflammatory cytokines
TNF-
(Fig. 4C) and IFN-
(Fig. 4D). These cytokines were
readily detected in culture supernatants of stimulated 2B4.11 or
2B4.hCD40 cells, but not Jurkat cells, which have been propagated
as an IL-2-producing human leukemia T cell line (24) and which
require overexpression of other proteins to induce TNF-
(37) and
IFN-
(38) secretion. Unlike IL-2, CD40 stimulation alone was
able to stimulate both production of TNF-
( p 0.02 when 6%
2B4.hCD40 cells were present, p 0.001 at 100%) and IFN-
( p 0.02 when 6% 2B4.hCD40 cells were present, p 0.0001 at
100%) in those cells stably expressing CD40. Similar to IL-2,
CD40 stimulation of T cells expressing CD40 was able to augment
both CD3 and CD3 plus CD28-mediated cytokine production.
The role of T cell CD40 in up-regulation of cell surface
molecules
CD40 signaling is known to up-regulate a number of cell surface
molecules (27), including CD11
(LFA-1), CD54 (ICAM-1), and
CD95 (Fas), all of which play significant roles in T cell activation
(39 41). We also evaluated the ability of CD40 to contribute to
CD80 (B7-1) and CD86 (B7-2) up-regulation, an important com-
ponent of activation by other CD40-expressing cells (42), as well
as CD25, a marker of T cell activation (43). We compared baseline
expression of these cell surface molecules on 2B4.11 and
2B4.hCD40 cells, as well as receptor-specific up-regulation of
these molecules postactivation (Fig. 5). Both 2B4.11 and
2B4.hCD40 expressed basal CD80 (Fig. 5C), CD86 (Fig. 5D), and
CD11
(LFA-1; Fig. 5E), with higher expression in 2B4.hCD40
cells. CD3 plus CD28 stimulation induced up-regulation of all sur-
face molecule tested in both 2B4.11 and 2B4.hCD40 cells (right
graph in each panel, p 0.01 for all groups). CD40 signals alone,
and in conjunction with CD3 or CD3 plus CD28 stimulation, in-
duced up-regulation of all surface molecules tested in 2B4.hCD40
cells, but not the nontransfected 2B4.11 cell line. CD40 signals
augmented CD3-mediated up-regulation of cell surface molecules
similar to CD28 stimulation in 2B4.hCD40 cells ( p 0.01 CD3 vs
CD3 plus CD28 or CD3 plus CD40 for all groups, no significant
difference in response between CD3 plus CD28 and CD3 plus
CD40 stimulation), with a maximal response achieved when triple
CD3 plus CD28 plus CD40 stimulation was given. Interestingly,
CD40 signaling induced a maximal enhancement of the costimu-
latory response in 2B4.hCD40 cells in conjunction with CD3 plus
CD28 to up-regulate CD25 expression (Fig. 5A).
The role of T cell CD40 in activation of cytokine signaling
pathways
Experiments presented above demonstrate that CD40 can act as a
costimulatory molecule to enhance CD3 and CD3 plus CD28-me-
diated T cell activation. NFAT, AP-1, and NF-
B are known to be
involved in the activation of several T cell proinflammatory cyto-
kine genes, including those encoding IL-2 (44 46), IFN-
(47,
48), and TNF-
(45, 49). We and others have demonstrated that
CD40-mediated activation of B lymphocytes involves AP-1 (50,
51) and NF-
B (31, 52) signaling pathways, and there is evidence
FIGURE 4. CD40 enhances CD3 and CD3 plus
CD28-mediated cytokine production in 2B4.hCD40
and J.hCD40 lines. 2B4.11 2B4.hCD40 (A, C, and D)
or Jurkat J.hCD40 (B) cells were stimulated with
plate-bound anti-CD3 anti-CD28 and/or anti-CD40
Abs (compared with medium (Med)/isotype control
(IC)). Culture supernatants were collected after 24
(TNF-
), 48 (IL-2), or 72 (IFN-
) h and assayed for
IL-2 (A and B), TNF-
(C), or IFN-
(D) by ELISA.
Data represent the mean SEM of triplicate samples
from two independent experiments. There was no dif-
ference between medium and isotype control samples.
675The Journal of Immunology
of CD40-mediated NFAT activation (51, 53). We asked if this is
also the case for T cell CD40.
CD3 or CD40 signals alone induced minimal NFAT reporter
gene activation compared with control stimuli in 2B4.hCD40 cells
(Fig. 6A), while CD28 in conjunction with CD3 signals induced an
increased response ( p 0.001) compared with medium/isotype
controls or CD3 single stimulation in 2B4.hCD40 cells (Fig. 6A;
p 0.0001 compared with medium/isotype controls or CD3 in
J.hCD40 cells, Fig. 6B). While CD40 signaling was able to en-
hance CD3-mediated NFAT activation ( p 0.001 in both
2B4.hCD40 and J.hCD40 cells), it was not to the same degree as
CD28-mediated enhancement. The greatest NFAT response was
achieved via engagement of all three receptors: CD3 plus CD28 plus
CD40 ( p 0.007 compared with CD3 plus CD28, p 0.001 com-
pared with CD3 plus CD40 in 2B4.hCD40 cells, p 0.017 compared
with CD3 plus CD28, p 0.003 compared with CD3 plus CD40 in
J.hCD40 cells). Although CD3 signals alone were able to trigger in-
creased NFAT activity in J.hCD40 cells ( p 0.001; Fig. 6B), CD28
and CD40 signals augmented this response in a manner similar to
their effects in 2B4.hCD40 cells.
We next investigated the ability of CD40 to signal via AP-1 in
T cells (Fig. 6). Unlike NFAT (Fig. 6), CD40 signals alone acti-
vated AP-1 2-fold over control stimuli in 2B4.hCD40 cells (Fig.
7A; p 0.005). While CD40 signals enhanced CD3-mediated
AP-1 activation in both cell lines ( p 0.001 in 2B4.hCD40 cells;
p 0.009 in J.hCD40 cells), CD40 was a less efficient costimu-
lator in this response than CD28 (CD3 vs CD3 plus CD28: p
0.007 in 2B4.hCD40 cells; p 0.005 in J.hCD40), and the
FIGURE 5. CD40 enhances CD3
and CD3 plus CD28-mediated cell
surface molecule up-regulation.
2B4.11 or 2B4.hCD40 cells were
stimulated with plate-bound anti-
CD3 anti-CD28 and/or anti-CD40
Abs (compared with isotype control)
for 48 –72 h, then analyzed by flow
cytometry to determine expression
levels of surface proteins. Histograms
(gray isotype control; black sur-
face molecule) represent baseline ex-
pression (stimulated with isotype con-
trol Ab), with corresponding median
channel fluorescence (MCF MCF
of surface molecule MCF of iso-
type control staining) represented in
the left graph of each panel. Changes
in MCF for each surface molecule
tested (poststimulation minus base-
line) is represented in the right graph
of each panel. A–F, Data represent the
mean SEM of MCF from two in-
dependent experiments.
676 SIGNALING BY CD40 IN T CELLS
maximal response was again achieved with simultaneous engage-
ment of CD3 plus CD28 plus CD40 ( p 0.02 compared with CD3
plus CD28, p 0.03 compared with CD3 plus CD40 in
2B4.hCD40 cells, p 0.01 compared with CD3 plus CD28, p
0.002 compared with CD3 plus CD40 in J.hCD40 cells). CD40
signals in B cells strongly activate phosphorylation of the AP-1
family member c-jun, via activation of JNK (54 –56). We evalu-
ated the ability of T cell CD40 to contribute to JNK activation
alone or in combination with CD3 or CD3 plus CD28. Strikingly,
in both 2B4.hCD40 and J.hCD40 cells (Fig. 8, A and B), only when
CD40 was engaged was strong phosphorylation of JNK observed,
with no phosphorylation seen via CD3 signaling and minimal
phosphorylation via CD3 plus CD28 signaling.
It is well established that CD40 signals lead to NF-
B activation in
B cells (52, 56–58) and we asked if this was also true for T cell CD40.
CD40 alone initiated a strong NF-
B response in 2B4.hCD40 (Fig.
9A, p 0.001) and J.hCD40 (Fig. 9B, p 0.001) cells, while CD3
evoked a minimal response compared with controls. CD3 plus CD40
activated NF-
B more than CD40 alone (CD40 vs CD3 plus CD40:
p 0.03 in 2B4.hCD40 cells, p 0.02 in J.hCD40 cells) and 2.5- to
5-fold greater than CD3 plus CD28 (CD3 plus CD28 vs CD3 plus
CD40: p 0.005 in 2B4.hCD40 cells; p 0.004 in J.hCD40 cells),
while the combination of CD3 plus CD28 plus CD40 signals provided
an additional 20% increase in NF-
B activation ( p 0.001 compared
with CD3 plus CD28, p 0.03 compared with CD3 plus CD40 in
2B4.hCD40 cells; p 0.01 compared with CD3 plus CD28, p 0.04
compared with CD3 plus CD40 in J.hCD40 cells).
Early events in NF-
B activation have been shown to include
phosphorylation and degradation of I
B
(59), and recent studies
suggest that an alternate pathway for activating NF-
B (NF-
B2),
in which p100 is processed to p52 and shuttled to the nucleus by
RelB, is also used by some TNFR family members, including
CD40 (56, 58). To test which NF-
B activation pathways are used
by CD40 in T cells, we assayed for I
B
phosphorylation and
degradation (NF-
B1; Fig. 10) or processing of p100 to p52 with
nuclear translocation of p52 and RelB (NF-
B2; Fig. 11). Using
densitometry to normalize I
B
values to the loading control actin
(Fig. 10B), CD40 engagement alone stimulated phosphorylation
and degradation of I
B
with up to 100% greater efficiency than
CD3 or CD3 plus CD28, with an even greater increase when CD40
signals were combined with CD3 or CD3 plus CD28. Similar re-
sults were seen in J.hCD40 cells (data not shown). With respect to
the NF-
B2 pathway, in both 2B4.hCD40 (Fig. 11) and J.hCD40
(data not shown), CD40 stimulation alone, but not CD3, resulted in
processing of p100 to p52 (Fig. 11, A and B) and translocation of
p52 and RelB to the nucleus (Fig. 11, C and D). Using densitom-
etry to normalize p100, p52, and RelB values to the loading control
actin (Fig. 11B, cytoplasmic fraction) or yy1 (Fig. 11D, nuclear
FIGURE 7. CD40 activates and enhances CD3 and CD3 plus CD28-
mediated AP-1 activation. 2B4.hCD40 (A) or J.hCD40 (B) cells were tran-
siently transfected with 7 AP-1-luciferase and Renilla-luciferase reporter
plasmids, rested on ice for 30 min, then stimulated for 24 h. Cells were
stimulated with anti-hCD40 or isotype control (IC) Abs or Dynal beads
armed with anti-CD3 or anti-CD3/anti-CD28 Abs. Relative luciferase ac-
tivity (AP-1:Renilla) of stimulus vs control cell groups was calculated as
the mean SEM of duplicate samples from two independent experiments.
There was no difference between medium and isotype control samples.
FIGURE 6. CD40 enhances CD3 and CD3 plus
CD28-mediated NFAT activation. 2B4.hCD40 (A)or
J.hCD40 (B) cells were transiently transfected with 4
NFAT-luciferase and Renilla-luciferase reporter plas-
mids, rested on ice for 30 min, then stimulated for 24 h.
Cells were stimulated with anti-hCD40 or isotype con-
trol (IC) Abs or Dynal beads armed with anti-CD3 or
anti-CD3 plus anti-CD28 Abs. Relative luciferase ac-
tivity (NFAT:Renilla) of stimulus vs control was cal-
culated as the mean SEM of duplicate samples from
three independent experiments. There was no difference
between medium and isotype control samples.
677The Journal of Immunology
fraction), we observed that CD40 and CD28 augmented CD3-me-
diated activation of NF-
B2 to a similar degree, whereas the com-
bination of CD3 plus CD28 plus CD40 gave maximal stimulation.
TRAF association of T cell CD40
CD40, like other members of the TNFR superfamily, relies on the
association of adaptor molecules, TRAFs, for downstream signal-
ing events, including activation of kinases and transcription fac-
tors, production of cytokines, up-regulation of surface molecules,
and various aspects of the humoral response (20). However, the
characteristics of TRAF association with CD40 have been shown
to differ between B cells, macrophages, dendritic cells, and epi-
thelial cells (19). It was thus important to determine CD40-TRAF
associations in T cells. In mouse B cells, we have previously dem-
onstrated that TRAFs 1, 2, 3, and 6 associate with either endoge-
nous mouse CD40 or transfected hCD40, following receptor liga-
tion (19, 60 62). We therefore compared the ability of TRAFs to
move into membrane lipid rafts (Fig. 12, A and B, left panel) and
associate with hCD40 (Fig. 12, A and B, right panel) in Brij sol-
uble (cytoplasmic) and insoluble (lipid raft) fractions in T cells.
TRAF2 and TRAF3 moved efficiently into the lipid raft fraction (left
panel) and associated with CD40 (right panel) in both 2B4.hCD40
(Fig. 12A) and J.hCD40 (Fig. 12B) cells, similar to their recruitment
in B cells (61). This was true if hCD40 was engaged by agonistic Abs
or by CD154, although more efficient movement (left panel) and re-
ceptor association (right panel) of TRAF1 and TRAF6 occurred when
hCD154 was the stimulus, as previously reported (63, 64).
Discussion
All B cells express CD40, and its functions on B cells have been
the subject of much study. Much less well appreciated is that ac-
tivated T cells can also express CD40, and the roles and mecha-
nisms of action of T cell-expressed CD40 still represent a signif-
icant knowledge gap. Various TNFR family members have been
shown to act as coregulatory molecules on T cells, some possibly
contributing to inflammatory disease (65). This role has been pre-
viously established for CD40 on B cells (66), and we demonstrate
here that this is also true for T cell CD40. We see a reproducible
increase in CD40 expression on T cells from mice with the chronic
inflammatory disorder CIA (Fig. 1 and Table I). Although the in-
crease in CD40
cells in CIA vs control mice is modest compared
with genetically predisposed autoimmune-prone strains of mice
(14, 17), it is consistent with the FACS profile of T cells from
normal mice (9, 17, 67). Strains of mice that are prone to sponta-
neous autoimmune disease have robust expression of CD40 by T
cells, whereas nondisease prone strains do not (14), so it is not
surprising that C57BL/6 mice do not have as robust proportions of
CD40
T cells as strains that develop autoimmune diseases rela
-
tively early in life. A normal mouse strain can induce functional
CD40 on T cells when encountering an Ag/danger signal (67), but
this level still does not reach that of mouse strains prone to spon-
taneous autoimmune disease, which have likely been receiving
chronic stimulation since early in life. However, findings presented
here show that T cell CD40 can contribute to enhanced T cell
activation even in individuals who do not have strong genetic pre-
disposition to autoimmunity.
Importantly, this CD40 is capable of T cell stimulation, indicat-
ing that it may have significant functional consequences (9, 14, 15,
67). Initially, this was a surprising finding. However, the Ag-spe-
cific precursor frequency for naive T cells is 1/1 10
5
(100
cells/spleen) (68, 69), a small population capable of significantly
expanding and producing a sufficient protective adaptive immune
response upon antigenic stimulation. This response ultimately
leads to the survival of 5% of the activated T cells (1 10
5
cells) to serve as a memory pool after the contraction phase (69).
Interestingly, in the CIA model, at 70 days postimmunization, we
can isolate a population of 3– 4 10
5
CD40
T cells per spleen
(Table I) that are capable of CD40-mediated activation of cytokine
FIGURE 8. CD40 activates and
enhances CD3 and CD3 plus CD28-
mediated JNK phosphorylation.
2B4.hCD40 (A) or J.hCD40 (B)
cells were rested at 37°C for 30 min,
then stimulated for 5–60 min with
anti-hCD40 or isotype control (IC)
Abs or Dynal beads armed with anti-
CD3 or anti-CD3 plus anti-CD28
Abs. Cells were pelleted, lysed, and
lysates analyzed by SDS-PAGE and
Western blot for phosphorylated (P.)
and total (T.) JNK. Similar results
were obtained in two independent
experiments.
FIGURE 9. CD40 activates and enhances CD3 and CD3 plus CD28-
mediated NF-
B activation. 2B4.hCD40 (A) or J.hCD40 (B) cells were
transiently transfected with 4 NF-
B-luciferase and Renilla-luciferase
reporter plasmids, rested for 15 min on ice, then stimulated for 6 h. Cells
were stimulated with anti-hCD40 or isotype control (IC) Abs or Dynal
beads armed with anti-CD3 or anti-CD3 plus anti-CD28 Abs. Relative
luciferase activity (NF-
B:Renilla) of stimulus vs control was calculated as
the mean SEM of duplicate samples from two independent experiments.
There was no difference between medium and isotype control samples.
678 SIGNALING BY CD40 IN T CELLS
production (Fig. 2), despite only representing 7% of the isolated
T cell population (Table I). This finding was recapitulated in ex-
periments whereby CD40
T cells were mixed with CD40
T
cells in a controlled fashion and a significant CD40-mediated re-
sponse was seen with as few as 6% CD40
T cells in culture
(Fig. 4).
Like other costimulatory TNFR family members expressed on T
cells (70), while CD40 itself cannot induce IL-2 production, it
augments the CD3 response and gives maximal stimulation to-
gether with CD3 plus CD28 signals (Figs. 2 and 4). This is true in
both T cells from CIA mice (Fig. 2) and in T cell lines expressing
CD40 (Fig. 4, A and B), even when transfected CD40
cells are
mixed with CD40
(untransfected) T cells to give a similar small
percentage of CD40
cells as that observed in CIA mice. Impor
-
tantly, anti-CD40 stimulation does not yield a positive cytokine
response in T cells lacking CD40, although their response to CD3
plus CD28 stimulation is similar to that of CD40
T cells (Figs. 2
and 4). Activation of the transcription factor NFAT is critical to
IL-2 production (44). Also consistent with a role as a costimulator,
CD40 itself cannot induce NFAT activation (Fig. 6), but can aug-
ment these responses to CD3 and CD3 plus CD28 ligation. It is
likely that the TCR complex provides calcium-mediated signaling
that is necessary, but not sufficient, for T cell activation and IL-2
production (71), while CD40 provides costimulatory signaling via
NF-
B and AP-1, necessary to activate NFAT and subsequent IL-2
production (44, 72, 73).
FIGURE 10. CD40 activates and enhances I
B
phosphorylation/degradation. A, 2B4.hCD40 cells were
stimulated with anti-hCD40 or isotype control (IC) Abs
or Dynal beads armed with anti-CD3 or anti-CD3 plus
anti-CD28 Abs. Cells were pelleted, lysed, and lysates
analyzed by SDS-PAGE and Western blot for phos-
phorylated and total I
B
. B, Density of bands as a
proportion of cells treated with medium (med) alone
and normalized to the value of the actin bands shown in
B. Similar results were obtained in two independent
experiments.
FIGURE 11. CD40 activates and enhances CD3 and
CD3 plus CD28-mediated nuclear translocation of p52
and RelB in the NF-
B2 pathway. 2B4.hCD40 (A–D)
cells were stimulated with anti-hCD40 or isotype con-
trol (IC) Abs or Dynal beads armed with anti-CD3 or
anti-CD3/anti-CD28 Abs. Cells were pelleted, cyto-
plasmic and nuclear fractions isolated, and lysates an-
alyzed by SDS-PAGE and Western blot for p100/p52
and RelB. Blots were stripped and reblotted for actin
(cytoplasmic fractions; A and B) or yy1 (nuclear frac-
tions; C and D) as loading controls. B, Density of bands
as a proportion of cells treated with medium (med)
alone and normalized to the value of the actin bands
shown in A. D, Density of bands as a proportion of cells
treated with medium (med) alone, normalized to the
value of the yy1 bands shown in C. Similar results were
obtained in two independent experiments.
679The Journal of Immunology
Both CD40 (74) and CD28 (75) contribute to cell activation via
association with lipid rafts. CD40 signals use TRAF adaptor mol-
ecules in B cells (61, 64), and we demonstrate in Fig. 12 that the
same is true for T cell CD40. TRAFs 1, 2, 3, and 6 bind CD40
within the Brij insoluble (raft) fraction upon engagement with ag-
onistic anti-CD40 Ab or membrane-bound CD154. This suggests
that TRAFs are a critical component of T cell CD40 signaling,
providing a key difference from CD28-mediated signaling. As in B
cells (50), T cell CD40 efficiently activates up-regulation of cell
surface molecules (Fig. 5), both canonical and noncanonical
NF-
B pathways (Figs. 10 and 11), AP-1 (Fig. 7), and the AP-1
activator JNK (Fig. 8). CD40 as a costimulatory molecule is as
effective as CD28 at signaling via AP-1 and severalfold more ef-
ficient at signaling via NF-
B, with maximal increase in both re-
sponses when stimulating T cells with agonists for CD3 plus CD28
plus CD40. This suggests that CD40 and CD28 may have different
molecular mechanisms leading to activation of AP-1 and NF-
B,
as has been proposed when comparing CD28 and other TNFR
family members on T cells (76). Importantly, it shows that CD40
can provide more powerful enhancement of NF-
B, a transcription
factor that induces cytokine genes with particular potency, than
other T cell costimulators. This is also seen in the activation of
JNK, which in the two T cell lines tested, was seen only when a
CD40 costimulus was included.
The above findings suggest that CD40 can increase the potency
and number of signaling pathways available to T cells that express
it. This is important when considering threshold requirements for
developing autoimmune disease and associated chronic inflamma-
tion. The presence of CD40-expressing T cells or autoantibodies is
not enough to develop autoimmunity. NOR mice, like their NOD
counterparts, have CD40-expressing T cells, yet do not develop
disease (17). Similarly, the presence of autoantibodies does not
necessarily indicate pathogenesis (77–79). Cooperation between
cell types and signals from the environment to the immune re-
sponse determine development of disease (80, 81).
Like many autoimmune diseases, including diabetes (82), ar-
thritis (83), and lupus (7), CIA requires cooperation between T and
B cells in its pathogenesis (84), and CD40 maybe a potent co-
stimulator in this process. CD40 expressed by B and T cells may
use similar molecular mechanisms, described above, to contribute
to the pathogenesis of inflammation in autoimmune disease. CD40
on both B cells (18, 85– 89) and T cells (this study) synergizes with
AgRs to enhance lymphocyte activation, cytokine production
(Figs. 2 and 4), and up-regulation of cell surface molecules (Fig.
5). CD40 signaling leads to isotype switching and autoantibody
production in B cells (90). In T cells, it has been demonstrated that
CD40 engagement leads to TCR revision within germinal centers
(67), skewing the T cell population further toward autoimmunity
(14 –16).
A proinflammatory cytokine environment is critical for reaching
the threshold of autoimmune disease development (65, 91). CD40
engagement in either T or B cells leads to TNF-
secretion, as
shown in Figs. 2C and 4C and (56, 92). We have previously re-
ported that CD40 signaling to B cells is partially mediated by
TNF-
binding to TNFR2 (56, 92) and hypothesize this also to be
true of T cell CD40, as it has been demonstrated that TNF-
acts
via TNFR2 to lower the threshold of T cell activation and IL-2
production (93–95). Taken together, the presence of increased
numbers of CD40
T cells in autoimmune mice and the demon
-
stration that CD40 can act as an effective costimulatory receptor on
T cells suggest that blockade of CD40-CD154 interactions can
abrogate the pathogenesis of autoimmune disease on several lev-
els. This has been demonstrated in the experimental autoimmune
encephalomyelitis model of multiple sclerosis, whereby disease
development is dependent on CD40 signaling, particularly in the
absence of CD28 (96). In the HgCl
2
-induced autoimmunity model,
CD28 is unable to overcome the lack of CD40 signaling to induce
disease (97).
CD40 may not just be a TCR costimulatory molecule on T cells
but may make these cells effective APC by up-regulating cell sur-
face molecules (Fig. 5). Increased levels of not only CD40, but
other costimulatory molecules such as CD80, CD86, ICAM-1, and
LFA-1, may also be pivotal in autoimmune disease development
(97). Although beyond the scope of this study, it would be inter-
esting to investigate the costimulatory capacity of activated CD40-
expressing T cells for CD40-negative T cells, as well as for other
cell types, including APC. It is quite possible that activation of
CD40 on T cells lowers the threshold of disease development but
is not sufficient for it to occur, requiring CD40 activation on APC
for cytokine production and on B cells for autoantibody secretion.
In this way, CD40 would prove central to autoimmune disease
development and pathogenesis, influencing not only the cognitive
activation of APC but also T cells. As seen in Fig. 2, results in
enhanced effector proinflammatory cytokine production that along
with CD40 activation on B cells would lead to autoantibody pro-
duction and, ultimately, chronic inflammation.
Disclosures
The authors have no financial conflict of interest.
References
1. Grewal, I. S., and R. A. Flavell. 1996. The role of CD40 ligand in costimulation
and T cell activation. Immunol. Rev. 153: 85–106.
2. Rodriguez-Pinto, D., and J. Moreno. 2005. B cells can prime naive CD4
T cells
in vivo in the absence of other professional antigen-presenting cells in a CD154-
CD40-dependent manner. Eur. J. Immunol. 35: 1097–1105.
3. Bishop, G. A., and B. S. Hostager. 2003. The CD40-CD154 interaction in B
cell-T cell liaisons. Cytokine Growth Factor Rev. 14: 297–309.
4. Balasa, B., T. Krahl, G. Patstone, J. Lee, R. Tisch, H. O. McDevitt, and
N. Sarvetnick. 1997. CD40 ligand-CD40 interactions are necessary for the initi-
ation of insulitis and diabetes in nonobese diabetic mice. J. Immunol. 159:
4620 4627.
5. Durie, F. H., R. A. Fava, T. M. Foy, A. Aruffo, J. A. Ledbetter, and R. J. Noelle.
1993. Prevention of collagen-induced arthritis with an antibody to gp39, the
ligand for CD40. Science 261: 1328 –1330.
FIGURE 12. CD40-mediated TRAF recruitment. 2B4.hCD40 (A)or
J.hCD40 (B) cells were stimulated for 15 min with anti-hCD40 Ab (vs
isotype control Ab; IC) or Hi5 insect cells expressing hCD154 (vs WT
baculovirus; WT). Detergent soluble (S) and insoluble (I) lysates were
prepared as described in Materials and Methods. Seventy percent of the
lysates were immunoprecipitated with protein G-Sepharose armed with
anti-hCD40. Lysates and immunoprecipitates were separated by SDS-
PAGE and analyzed by Western blot. Similar results were obtained in two
identical experiments.
680 SIGNALING BY CD40 IN T CELLS
6. Toubi, E., and Y. Shoenfeld. 2004. The role of CD40-CD154 interactions in
autoimmunity and the benefit of disrupting this pathway. Autoimmunity 37:
457– 464.
7. Datta, S. K. 1998. Production of pathogenic antibodies: cognate interactions be-
tween autoimmune T and B cells. Lupus 7: 591–596.
8. Kyburz, D., M. Corr, D. C. Brinson, A. Von Damm, H. Tighe, and D. A. Carson.
1999. Human rheumatoid factor production is dependent on CD40 signaling and
autoantigen. J. Immunol. 163: 3116 –3122.
9. Fanslow, W. C., K. N. Clifford, M. Seaman, M. R. Alderson, M. K. Spriggs,
R. J. Armitage, and F. Ramsdell. 1994. Recombinant CD40 ligand exerts potent
biologic effects on T cells. J. Immunol. 152: 4262– 4269.
10. Bourgeois, C., B. Rocha, and C. Tanchot. 2002. A role for CD40 expression on
CD8
T cells in the generation of CD8
T cell memory. Science 297:
2060 –2063.
11. Koschella, M., D. Voehringer, and H. Pircher. 2004. CD40 ligation in vivo in-
duces bystander proliferation of memory phenotype CD8 T cells. J. Immunol.
172: 4804 4811.
12. Lee, B. O., L. Hartson, and T. D. Randall. 2003. CD40-deficient, influenza-
specific CD8 memory T cells develop and function normally in a CD40-sufficient
environment. J. Exp. Med. 198: 1759 –1764.
13. Sun, J. C., and M. J. Bevan. 2004. Cutting edge: long-lived CD8 memory and
protective immunity in the absence of CD40 expression on CD8 T cells. J. Im-
munol. 172: 3385–3389.
14. Wagner, D. H., Jr., E. Newell, R. J. Sanderson, J. H. Freed, and M. K. Newell.
1999. Increased expression of CD40 on thymocytes and peripheral T cells in
autoimmunity: a mechanism for acquiring changes in the peripheral T cell re-
ceptor repertoire. Int. J. Mol. Med. 4: 231–242.
15. Wagner, D. H., Jr., G. Vaitaitis, R. Sanderson, M. Poulin, C. Dobbs, and
K. Haskins. 2002. Expression of CD40 identifies a unique pathogenic T cell
population in type 1 diabetes. Proc. Natl. Acad. Sci. USA 99: 3782–3787.
16. Vaitaitis, G. M., M. Poulin, R. J. Sanderson, K. Haskins, and D. H. Wagner Jr.
2003. Cutting edge: CD40-induced expression of recombination activating gene
(RAG) 1 and RAG2: a mechanism for the generation of autoaggressive T cells in
the periphery. J. Immunol. 170: 3455–3459.
17. Waid, D. M., G. M. Vaitaitis, and D. H. Wagner, Jr. 2004. Peripheral
CD4
lo
CD40
auto-aggressive T cell expansion during insulin-dependent diabetes
mellitus. Eur. J. Immunol. 34: 1488 –1497.
18. Haxhinasto, S. A., B. S. Hostager, and G. A. Bishop. 2002. Cutting edge: mo-
lecular mechanisms of synergy between CD40 and the B cell antigen receptor:
role for TNF receptor-associated factor 2 in receptor interaction. J. Immunol. 169:
1145–1149.
19. Bishop, G. A., and B. S. Hostager. 2001. Molecular mechanisms of CD40 sig-
naling. Arch. Immunol. Ther. Exp. 49: 129 –137.
20. Bishop, G. A., and B. S. Hostager. 2001. Signaling by CD40 and its mimics in
B cell activation. Immunol. Res. 24: 97–109.
21. Croft, M. 2003. Co-stimulatory members of the TNFR family: keys to effective
T cell immunity? Nat. Rev. Immunol. 3: 609 620.
22. Watts, T. H. 2005. TNF/TNFR family members in costimulation of T cell re-
sponses. Ann. Rev. Immunol. 23: 23– 68.
23. Ashwell, J. D., R. E. Cunningham, P. D. Noguchi, and D. Hernandez. 1987. Cell
growth cycle block of T cell hybridomas upon activation with antigen. J. Exp.
Med. 165: 173–194.
24. Gillis, S., and J. Watson. 1980. Biochemical and biological characterization of
lymphocyte regulatory molecules. V. Identification of an interleukin 2-producing
human leukemia T cell line. J. Exp. Med. 152: 1709 –1719.
25. Warren, W. D., and M. T. Berton. 1995. Induction of germ-line
1 and Ig gene
expression in murine B cells—IL-4 and the CD40 ligand-CD40 interaction pro-
vide distinct but synergistic signals. J. Immunol. 155: 5637–5646.
26. Bishop, G. A., W. D. Warren, and M. T. Berton. 1995. Signaling via major
histocompatibility complex class II molecules and antigen receptors enhances the
B cell response to gp39/CD40 ligand. Eur. J. Immunol. 25: 1230 –1238.
27. Hostager, B. S., Y. Hsing, D. E. Harms, and G. A. Bishop. 1996. Different
CD40-mediated signaling events require distinct CD40 structural features. J. Im-
munol. 157: 1047–1053.
28. Bishop, G. A., and J. A. Frelinger. 1989. Haplotype-specific differences in sig-
naling by transfected class II molecules to a Ly-1
B cell clone. Proc. Natl. Acad.
Sci. USA 86: 5933–5937.
29. Campbell, I. K., J. A. Hamilton, and I. P. Wicks. 2000. Collagen-induced arthritis
in C57BL/6 (H-2
b
) mice: new insights into an important disease model of rheu
-
matoid arthritis. Eur. J. Immunol. 30: 1568–1575.
30. Rosloniec, E. F., A. H. Kang, L. K. Myers, and M. A. Cremer. 1997. Collagen-
Induced Arthritis. Wiley & Sons, New York.
31. Berberich, I., G. L. Shu, and E. A. Clark. 1994. Cross-linking CD40 on B cells
rapidly activates nuclear factor
B. J. Immunol. 153: 4357– 4366.
32. Schreiber, E., P. Matthias, M. M. Muller, and W. Schaffner. 1989. Rapid detec-
tion of octamer binding proteins with “mini-extracts,” prepared from a small
number of cells. Nucleic Acids Res. 17: 6419.
33. Moore, C. R., and G. A. Bishop. 2005. Differential regulation of CD40-mediated
TNF receptor-associated factor degradation in B lymphocytes. J. Immunol. 175:
3780 –3789.
34. Xie, P., B. S. Hostager, and G. A. Bishop. 2004. Requirement for TRAF3 in
signaling by LMP1 but not CD40 in B lymphocytes. J. Exp. Med. 199: 661– 671.
35. Myers, L. K., E. F. Rosloniec, M. A. Cremer, and A. H. Kang. 1997. Collagen-
induced arthritis, an animal model of autoimmunity. Life Sci. 61: 1861–1878.
36. Schulze-Koops, H., and J. R. Kalden. 2001. The balance of Th1/Th2 cytokines in
rheumatoid arthritis. Best Pract. Res. Clin. Rheumatol. 15: 677– 691.
37. Ballester, A., A. Velasco, R. Tobena, and S. Alemany. 1998. Cot kinase activates
tumor necrosis factor
gene expression in a cyclosporin A-resistant manner.
J. Biol. Chem. 273: 14099 –14106.
38. Kashiwakura, J., N. Suzuki, H. Nagafuchi, M. Takeno, Y. Takeba,
Y. Shimoyama, and T. Sakane. 1999. Txk, a nonreceptor tyrosine kinase of the
Tec family, is expressed in T helper type 1 cells and regulates interferon
pro-
duction in human T lymphocytes. J. Exp. Med. 190: 1147–1154.
39. Chung, D. H., K. C. Jung, W. S. Park, I. S. Lee, W. J. Choi, C. J. Kim, S. H. Park,
and Y. Bae. 2000. Costimulatory effect of Fas in mouse T lymphocytes. Mol. Cell
10: 642– 646.
40. Tibbetts, S. A., C. Chirathaworn, M. Nakashima, D. S. Jois, T. J. Siahaan,
M. A. Chan, and S. H. Benedict. 1999. Peptides derived from ICAM-1 and
LFA-1 modulate T cell adhesion and immune function in a mixed lymphocyte
culture. Transplantation 68: 685–692.
41. Moulian, N., J. Bidault, C. Planche, and S. Berrih-Aknin. 1998. Two signaling
pathways can increase fas expression in human thymocytes. Blood 92:
1297–1307.
42. Bhatia, S., M. Edidin, S. C. Almo, and S. G. Nathenson. 2006. B7-1 and B7-2:
similar costimulatory ligands with different biochemical, oligomeric and signal-
ing properties. Immunol. Lett. 104: 70 –75.
43. Adler, S. H., E. Chiffoleau, L. Xu, N. M. Dalton, J. M. Burg, A. D. Wells,
M. S. Wolfe, L. A. Turka, and W. S. Pear. 2003. Notch signaling augments T cell
responsiveness by enhancing CD25 expression. J. Immunol. 171: 2896 –2903.
44. Cianferoni, A., M. Massaad, S. Feske, M. A. de la Fuente, L. Gallego,
N. Ramesh, and R. S. Geha. 2005. Defective nuclear translocation of nuclear
factor of activated T cells and extracellular signal-regulated kinase underlies de-
ficient IL-2 gene expression in Wiskott-Aldrich syndrome. J. Allergy Clin. Im-
munol. 116: 1364 –1371.
45. Palanki, M. S. 2002. Inhibitors of AP-1 and NF-
B mediated transcriptional
activation: therapeutic potential in autoimmune diseases and structural diversity.
Curr. Med. Chem. 9: 219 –227.
46. Jain, J., C. Loh, and A. Rao. 1995. Transcriptional regulation of the IL-2 gene.
Curr. Opin. Immunol. 7: 333–342.
47. Kaminuma, O., C. Elly, Y. Tanaka, A. Mori, Y. C. Liu, A. Altman, and
S. Miyatake. 2002. Vav-induced activation of the human IFN-
gene promoter is
mediated by upregulation of AP-1 activity. FEBS Lett. 514: 153–158.
48. Sica, A., L. Dorman, V. Viggiano, M. Cippitelli, P. Ghosh, N. Rice, and
H. A. Young. 1997. Interaction of NF-
B and NFAT with the interferon
pro-
moter. J. Biol. Chem. 272: 30412–30420.
49. Li-Weber, M., M. K. Treiber, M. Giaisi, K. Palfi, N. Stephan, S. Parg, and
P. H. Krammer. 2005. Ultraviolet irradiation suppresses T cell activation via
blocking TCR-mediated ERK and NF-
B signaling pathways. J. Immunol. 175:
2132–2143.
50. Baccam, M., S. Y. Woo, C. Vinson, and G. A. Bishop. 2003. CD40-mediated
transcriptional regulation of the IL-6 gene in B lymphocytes: involvement of
NF-
B, AP-1, and C/EBP. J. Immunol. 170: 3099 –3108.
51. Francis, D. A., J. G. Karras, X. Y. Ke, R. Sen, and T. L. Rothstein. 1995. In-
duction of the transcription factors NF-
B, AP-1 and NF-AT during B cell stim-
ulation through the CD40 receptor. Int. Immunol. 7: 151–161.
52. Hsing, Y., and G. A. Bishop. 1999. Requirement for nuclear factor
B activation
by a distinct subset of CD40-mediated effector functions in B lymphocytes. J. Im-
munol. 162: 2804 –2811.
53. Antony, P., J. B. Petro, G. Carlesso, N. P. Shinners, J. Lowe, and W. N. Khan.
2003. B cell receptor directs the activation of NFAT and NF-
B via distinct
molecular mechanisms. Exp. Cell Res. 291: 11–24.
54. Brown, K. D., B. S. Hostager, and G. A. Bishop. 2001. Differential signaling and
tumor necrosis factor receptor-associated factor (TRAF) degradation mediated by
CD40 and the Epstein-Barr virus oncoprotein latent membrane protein 1 (LMP1).
J. Exp. Med. 193: 943–954.
55. Li, Y. Y., M. Baccam, S. B. Waters, J. E. Pessin, G. A. Bishop, and
G. A. Koretzky. 1996. CD40 ligation results in protein kinase C-independent
activation of ERK and JNK in resting murine splenic B cells. J. Immunol. 157:
1440 –1447.
56. Munroe, M. E., and G. A. Bishop. 2004. Role of tumor necrosis factor (TNF)
receptor-associated factor 2 (TRAF2) in distinct and overlapping CD40 and TNF
receptor 2/CD120b-mediated B lymphocyte activation. J. Biol. Chem. 279:
53222–53231.
57. Hsing, Y., B. S. Hostager, and G. A. Bishop. 1997. Characterization of CD40
signaling determinants regulating nuclear factor
B activation in B lymphocytes.
J. Immunol. 159: 48984906.
58. Coope, H. J., P. G. Atkinson, B. Huhse, M. Belich, J. Janzen, M. J. Holman,
G. G. Klaus, L. H. Johnston, and S. C. Ley. 2002. CD40 regulates the processing
of NF-
B2 p100 to p52. EMBO J. 21: 5375–5385.
59. Karin, M., and Y. Ben-Neriah. 2000. Phosphorylation meets ubiquitination: the
control of NF-
B activity. Ann. Rev. Immunol. 18: 621– 663.
60. Bishop, G. A., B. S. Hostager, and K. D. Brown. 2002. Mechanisms of TNF
receptor-associated factor (TRAF) regulation in B lymphocytes. J. Leukocyte
Biol. 72: 19 –23.
61. Hostager, B. S., I. M. Catlett, and G. A. Bishop. 2000. Recruitment of CD40 and
tumor necrosis factor receptor-associated factors 2 and 3 to membrane microdo-
mains during CD40 signaling. J. Biol. Chem. 275: 15392–15398.
62. Jalukar, S. V., B. S. Hostager, and G. A. Bishop. 2000. Characterization of the
roles of TNF receptor-associated factor 6 in CD40-mediated B lymphocyte ef-
fector functions. J. Immunol. 164: 623– 630.
63. Hostager, B. S., S. A. Haxhinasto, S. L. Rowland, and G. A. Bishop. 2003. Tumor
necrosis factor receptor-associated factor 2 (TRAF2)-deficient B lymphocytes
681The Journal of Immunology
reveal novel roles for TRAF2 in CD40 signaling. J. Biol. Chem. 278:
45382– 453890.
64. Xie, P., B. S. Hostager, M. E. Munroe, C. R. Moore, and G. A. Bishop. 2006.
Cooperation between TNF receptor-associated factors 1 and 2 in CD40 signaling.
J. Immunol. 176: 5388–5400.
65. Kwon, B., B. S. Kim, H. R. Cho, J. E. Park, and B. S. Kwon. 2003. Involvement
of tumor necrosis factor receptor superfamily (TNFRSF) members in the patho-
genesis of inflammatory diseases. Exp. Mol. Med. 35: 8 –16.
66. Lorenz, H. M., M. Herrmann, and J. R. Kalden. 2001. The pathogenesis of au-
toimmune diseases. Scand. J. Clin. Lab. Invest. 235: 16 –26.
67. Cooper, C. J., G. L. Turk, M. Sun, A. G. Farr, and P. J. Fink. 2004. Cutting edge:
TCR revision occurs in germinal centers. J. Immunol. 173: 6532– 6536.
68. Whitmire, J. K., N. Benning, and J. L. Whitton. 2006. Precursor frequency, non-
linear proliferation, and functional maturation of virus-specific CD4
T cells.
J. Immunol. 176: 3028–3036.
69. Blattman, J. N., R. Antia, D. J. Sourdive, X. Wang, S. M. Kaech,
K. Murali-Krishna, J. D. Altman, and R. Ahmed. 2002. Estimating the precursor
frequency of naive antigen-specific CD8 T cells. J. Exp. Med. 195: 657– 664.
70. Saoulli, K., S. Y. Lee, J. L. Cannons, W. C. Yeh, A. Santana, M. D. Goldstein,
N. Bangia, M. A. DeBenedette, T. W. Mak, Y. Choi, and T. H. Watts. 1998.
CD28-independent, TRAF2-dependent costimulation of resting T cells by 4-1BB
ligand. J. Exp. Med. 187: 1849 –1862.
71. Gupta, S. 1989. Mechanisms of transmembrane signalling in human T cell acti-
vation. Mol. Cell. Biochem. 91: 45–50.
72. Flescher, E., J. A. Ledbetter, N. Ogawa, N. Vela-Roch, D. Fossum, H. Dang, and
N. Talal. 1995. Induction of transcription factors in human T lymphocytes by
aspirin-like drugs. Cell. Immunol. 160: 232–239.
73. Isakov, N., and A. Altman. 2002. Protein kinase C
in T cell activation. Ann. Rev.
Immunol. 20: 761–794.
74. Kaykas, A., K. Worringer, and B. Sugden. 2001. CD40 and LMP-1 both signal
from lipid rafts but LMP-1 assembles a distinct, more efficient signaling complex.
EMBO J. 20: 2641–2654.
75. Kovacs, B., R. V. Parry, Z. Ma, E. Fan, D. K. Shivers, B. A. Freiberg,
A. K. Thomas, R. Rutherford, C. A. Rumbley, J. L. Riley, and T. H. Finkel. 2005.
Ligation of CD28 by its natural ligand CD86 in the absence of TCR stimulation
induces lipid raft polarization in human CD4 T cells. J. Immunol. 175:
7848 –7854.
76. Watts, T. H., and M. A. DeBenedette. 1999. T cell co-stimulatory molecules other
than CD28. Curr. Opinion. Immunol. 11: 286–293.
77. Nowak, U. M., and M. M. Newkirk. 2005. Rheumatoid factors: good or bad for
you? Int. Arch. Allergy Immunol. 138: 180 –188.
78. Schlosser, M., K. Koczwara, H. Kenk, M. Strebelow, I. Rjasanowski,
R. Wassmuth, P. Achenbach, A. G. Ziegler, and E. Bonifacio. 2005. In insulin-
autoantibody-positive children from the general population, antibody affinity
identifies those at high and low risk. Diabetologia 48: 1830 –1832.
79. Quintana, F. J., and I. R. Cohen. 2004. The natural autoantibody repertoire and
autoimmune disease. Biomed. Pharmacother. 58: 276 –281.
80. Matsui, K., S. Jodo, S. Xiao, and S. T. Ju. 1999. Hypothesis: a recurrent, mod-
erate activation fosters systemic autoimmunity—the apoptotic roles of TCR, IL-2
and Fas ligand. J. Biomed. Sci. 6: 306 –313.
81. Shepshelovich, D., and Y. Shoenfeld. 2006. Prediction and prevention of auto-
immune diseases: additional aspects of the mosaic of autoimmunity. Lupus 15:
183–190.
82. Noorchashm, H., S. A. Greeley, and A. Naji. 2003. The role of t/b lymphocyte
collaboration in the regulation of autoimmune and alloimmune responses. Immu-
nol. Res. 27: 443–450.
83. Korganow, A. S., H. Ji, S. Mangialaio, V. Duchatelle, R. Pelanda, T. Martin,
C. Degott, H. Kikutani, K. Rajewsky, J. L. Pasquali, et al. 1999. From systemic
T cell self-reactivity to organ-specific autoimmune disease via immunoglobulins.
Immunity 10: 451– 461.
84. Brand, D. D., A. H. Kang, and E. F. Rosloniec. 2003. Immunopathogenesis of
collagen arthritis. Springer Semin. Immunopathol. 25: 3–18.
85. Haxhinasto, S. A., and G. A. Bishop. 2003. Synergistic B cell activation by CD40
and the B cell antigen receptor: role of BCR-mediated kinase activation and
TRAF regulation. J. Biol. Chem. 279: 2575–2582.
86. Haxhinasto, S. A., and G. A. Bishop. 2003. A novel interaction between protein
kinase D and TNF receptor-associated factor molecules regulates B cell receptor-
CD40 synergy. J. Immunol. 171: 4655– 4662.
87. Mizuno, T., and T. L. Rothstein. 2003. Cutting edge: CD40 engagement elimi-
nates the need for Bruton’s tyrosine kinase in B cell receptor signaling for NF-
B. J. Immunol. 170: 2806 –2810.
88. Mizuno, T., and T. L. Rothstein. 2005. B cell receptor (BCR) cross-talk: CD40
engagement creates an alternate pathway for BCR signaling that activates I
B
kinase/I
B
/NF-
B without the need for PI3K and phospholipase C
. J. Immu-
nol. 174: 6062– 6070.
89. Mizuno, T., and T. L. Rothstein. 2005. B cell receptor (BCR) cross-talk: CD40
engagement enhances BCR-induced ERK activation. J. Immunol. 174:
3369 –3376.
90. Ollila, J., and M. Vihinen. 2005. B cells. Int. J. Biochem. Cell Biol. 37: 518 –523.
91. Kollias, G., and D. Kontoyiannis. 2002. Role of TNF/TNFR in autoimmunity:
specific TNF receptor blockade may be advantageous to anti-TNF treatments.
Cytokine Growth Factor Rev. 13: 315–321.
92. Hostager, B. S., and G. A. Bishop. 2002. Role of TNF receptor-associated factor
2 in the activation of IgM secretion by CD40 and CD120b. J. Immunol. 168:
3318 –3322.
93. Kim, E. Y., and H. S. Teh. 2004. Critical role of TNF receptor type-2 (p75) as a
costimulator for IL-2 induction and T cell survival: a functional link to CD28.
J. Immunol. 173: 45004509.
94. Kim, E. Y., and H. S. Teh. 2001. TNF type 2 receptor (p75) lowers the threshold
of T cell activation. J. Immunol. 167: 6812– 6820.
95. Aspalter, R. M., M. M. Eibl, and H. M. Wolf. 2003. Regulation of TCR-mediated
T cell activation by TNF-RII. J. Leukocyte Biol. 74: 572–582.
96. Girvin, A. M., M. C. Dal Canto, and S. D. Miller. 2002. CD40/CD40L interaction
is essential for the induction of EAE in the absence of CD28-mediated co-stim-
ulation. J. Autoimmun. 18: 83–94.
97. Pollard, K. M., M. Arnush, P. Hultman, and D. H. Kono. 2004. Costimulation
requirements of induced murine systemic autoimmune disease. J. Immunol. 173:
5880 –5887.
682 SIGNALING BY CD40 IN T CELLS
... For example, lck (lymphocyte-specific protein tyrosine kinase) and zap-70 (zeta-chain-associated protein kinase 70), crucial T cell receptor signaling molecules, were significantly down-regulated in infected WL stickleback (Table 2). Cd40, a costimulatory protein that is expressed by antigen presenting cells (APCs) and required for T and B cell activation [56,57] was also down-regulated, which is in line with previous reports in lab raised stickleback [13]. We also observed a significant down-regulation of sh2d1ab (sh2 domain containing 1A duplicate b), expressed in T cells, natural killer cells, and some B cells, where it modulates signal transduction pathways [58]. ...
Article
Full-text available
Wild organisms are regularly exposed to a wide range of parasites, requiring the management of an effective immune response while avoiding immunopathology. Currently, our knowledge of immunoparasitology primarily derives from controlled laboratory studies, neglecting the genetic and environmental diversity that contribute to immune phenotypes observed in wild populations. To gain insight into the immunologic variability in natural settings, we examined differences in immune gene expression of two Alaskan stickleback (Gasterosteus aculeatus) populations with varying susceptibility to infection by the cestode Schistocephalus solidus. Between these two populations, we found distinct immune gene expression patterns at the population level in response to infection with fish from the high-infection population displaying signs of parasite-driven immune manipulation. Further, we found significant differences in baseline immune gene profiles between the populations, with uninfected low-infection population fish showing signatures of inflammation compared to uninfected high-infection population fish. These results shed light on divergent responses of wild populations to the same parasite, providing valuable insights into host-parasite interactions in natural ecosystems.
... 41 Moreover, CD40 may induce the production of IFN-γ, TNF-α, and IL-2, consistent with the response profile in alpha1-oleate treated patients. 42 Anti-TNF therapy of autoimmune disorders has been associated with an increased risk for malignancies including bladder cancer. 55 Blocking of IL-6 signaling has been discussed as a potential therapeutic strategy for cancers characterized by pathological IL-6 overproduction. ...
Article
Full-text available
Background The molecular content of urine is defined by filtration in the kidneys and by local release from tissues lining the urinary tract. Pathological processes and different therapies change the molecular composition of urine and a variety of markers have been analyzed in patients with bladder cancer. The response to BCG immunotherapy and chemotherapy has been extensively studied and elevated urine concentrations of IL‐1RA, IFN‐α, IFN‐γ TNF‐α, and IL‐17 have been associated with improved outcome. Methods In this study, the host response to intravesical alpha 1‐oleate treatment was characterized in patients with non‐muscle invasive bladder cancer by proteomic and transcriptomic analysis. Results Proteomic profiling detected a significant increase in multiple cytokines in the treatment group compared to placebo. The innate immune response was strongly activated, including IL‐1RA and pro‐inflammatory cytokines in the IL‐1 family (IL‐1α, IL‐1β, IL‐33), chemokines (MIP‐1α, IL‐8), and interferons (IFN‐α2, IFN‐γ). Adaptive immune mediators included IL‐12, Granzyme B, CD40, PD‐L1, and IL‐17D, suggesting broad effects of alpha 1‐oleate treatment on the tumor tissues. Conclusions The cytokine response profile in alpha 1‐oleate treated patients was similar to that reported in BCG treated patients, suggesting a significant overlap. A reduction in protein levels at the end of treatment coincided with inhibition of cancer‐related gene expression in tissue biopsies, consistent with a positive treatment effect. Thus, in addition to killing tumor cells and inducing cell detachment, alpha 1‐oleate is shown to activate a broad immune response with a protective potential.
... In light of previous reports that activated CD8 + cells express CD40, and that CD40 ligation may provide a direct costimulatory signal independent of the wellknown effects of CD40L on antigen-presenting cells (APCs), [31][32][33] we hypothesized that the contact-dependent effect of CD4 + cells on CD8 + cells might be mediated through CD40L-CD40 interactions. We also hypothesized that given the improved outcomes reported in patients with higher levels of CD27 + cells, 6 that CD70-CD27 interactions may also contribute. ...
Article
Full-text available
Background Cell culture conditions during manufacturing can impact the clinical efficacy of chimeric antigen receptor (CAR) T cell products. Production methods have not been standardized because the optimal approach remains unknown. Separate CD4⁺ and CD8⁺ cultures offer a potential advantage but complicate manufacturing and may affect cell expansion and function. In a phase 1/2 clinical trial, we observed poor expansion of separate CD8⁺ cell cultures and hypothesized that coculture of CD4⁺ cells and CD8⁺ cells at a defined ratio at culture initiation would enhance CD8⁺ cell expansion and simplify manufacturing. Methods We generated CAR T cells either as separate CD4⁺ and CD8⁺ cells, or as combined cultures mixed in defined CD4:CD8 ratios at culture initiation. We assessed CAR T cell expansion, phenotype, function, gene expression, and in vivo activity of CAR T cells and compared these between separately expanded or mixed CAR T cell cultures. Results We found that the coculture of CD8⁺ CAR T cells with CD4⁺ cells markedly improves CD8⁺ cell expansion, and further discovered that CD8⁺ cells cultured in isolation exhibit a hypofunctional phenotype and transcriptional signature compared with those in mixed cultures with CD4⁺ cells. Cocultured CAR T cells also confer superior antitumor activity in vivo compared with separately expanded cells. The positive impact of CD4⁺ cells on CD8⁺ cells was mediated through both cytokines and direct cell contact, including CD40L-CD40 and CD70-CD27 interactions. Conclusions Our data indicate that CD4⁺ cell help during cell culture maintains robust CD8⁺ CAR T cell function, with implications for clinical cell manufacturing.
... The first marker we identified was CD40 in alloimmunized patients. This result was expected, as CD40 is a prosurvival molecule with a costimulatory function for activated CD4 + TLs (52,53). High levels of CD40 are suggestive of an active phenotype favoring immune responses, such as alloimmunization (54). ...
Article
Full-text available
Introduction Acute myeloid leukemia (AML) is one of the commonest hematologic disorders. Due to the high frequency of disease- or treatment-related thrombocytopenia, AML requires treatment with multiple platelet transfusions, which can trigger a humoral response directed against platelets. Some, but not all, AML patients develop an anti-HLA immune response after multiple transfusions. We therefore hypothesized that different immune activation profiles might be associated with anti-HLA alloimmunization status. Methods We tested this hypothesis, by analyzing CD4+ T lymphocyte (TL) subsets and their immune control molecules in flow cytometry and single-cell multi-omics. Results A comparison of immunological status between anti-HLA alloimmunized and non-alloimmunized AML patients identified differences in the phenotype and function of CD4+ TLs. CD4+ TLs from alloimmunized patients displayed features of immune activation, with higher levels of CD40 and OX40 than the cells of healthy donors. However, the most notable differences were observed in non-alloimmunized patients. These patients had lower levels of CD40 and OX40 than alloimmunized patients and higher levels of PD1. Moreover, the Treg compartment of non-alloimmunized patients was larger and more functional than that in alloimmunized patients. These results were supported by a multi-omics analysis of immune response molecules in conventional CD4+ TLs, Tfh circulating cells, and Tregs. Discussion Our results thus reveal divergent CD4+ TL characteristics correlated with anti-HLA alloimmunization status in transfused AML patients. These differences, characterizing CD4+ TLs independently of any specific antigen, should be taken into account when considering the immune responses of patients to infections, vaccinations, or transplantations.
... When comparing our results with those from previous transcriptomic studies, we observed moderate overlap between expression patterns (Table 3). For example, we found cd40, a costimulatory protein that is expressed by antigen-presenting cells and required for T and B cell activation (Grewal and Flavell, 1996;Munroe and Bishop, 2007) to be down-regulated, which is in line with previous reports (Lohman et al., 2017a). Further, we found the previously discussed anxa2a to be up-regulated, similar to previous work by Fuess and colleagues (Fuess et al., 2021c) and Lohman and colleagues (Lohman et al., 2017a). ...
Preprint
Full-text available
Helminth parasites pose a significant threat to host survival and reproductive success, imposing strong selection pressure on hosts to evolve countermessures (e.g., immune responses and behavioral changes). To gain insights into the underlying mechanisms of host-parasite co-evolution, we examined differences in gene expression in immune tissues of two Alaskan stickleback ( Gasterosteus aculeatus ) populations with varying susceptibility to infection by the cestode Schistocephalus solidus . Our analyses revealed distinct patterns of immune gene expression at the population-level in response to infection. Infected fish from the high infection population displayed signs of immune manipulation by the parasite, whereas this phenomenon was absent in fish from the low infection population. Notably, we found significant differences in immune gene expression between the populations, with uninfected Rocky Lake fish showing up-regulation of innate immune genes associated with inflammation compared to uninfected Walby Lake fish. These findings highlight the divergent evolutionary paths taken by different stickleback populations in their response to the same parasite.
... It is becoming clearer that there are many interactions that can be ascribed to this molecule that depend on the inflammatory state and on the cell type being studied. The quest has long been to inhibit CD40 signals since it is well understood that CD40 is intimately involved in driving autoimmune disease (2)(3)(4)(31)(32)(33)(34)(35). ...
Article
Full-text available
CD40-signaling has long been a target in autoimmunity. Attempts to block signaling between CD40 and CD154 during clinical trials using monoclonal antibodies suffered severe adverse events. Previously, we developed a peptide, KGYY15, that targets CD40 and, in preclinical trials, prevents Type 1 Diabetes in >90% of cases and reverses new onset hyperglycemia in 56% of cases. It did so by establishing normal effector T cell levels rather than ablating the cells and causing immunosuppression. However, the relationship between KGYY15 and other elements of the complex signaling network of CD40 is not clear. Studying interactions between proteins from autoimmune and non-autoimmune mice, we demonstrate interactions between CD40 and integrin CD11a/CD18, which complicates the understanding of the inflammatory nexus and how to prevent auto-inflammation. In addition to interacting with CD40, KGYY15 interacts with the integrins CD11a/CD18 and CD11b/CD18. We argue that modulation of CD40-CD154 signaling may be more advantageous than complete inhibition because it may preserve normal immunity to pathogens.
... The CD40/CD40L interaction mediates a wide range of cellular activities. The engagement of CD40 on the surface of Antigen Presenting Cell induces the production of various cytokines and facilitates them to mature and attain the necessary characteristics required to initiate differentiation, proliferation, and activation of T cells [132][133][134][135][136]. CD40 is expressed in various malignancies such as colon cancer, breast cancer, ovarian cancer, and lung cancer. ...
Article
Objective Recent scientific advances have expanded insight into the immune system and its response to malignant cells. In the past few years, immunotherapy has attained a hallmark for cancer treatment, especially for patients suffering from the advanced-stage disease. Modulating the immune system by blocking various immune checkpoint receptor proteins through monoclonal antibodies has improved cancer patients' survival rates. Methods The scope of this review spans from 1985 to the present day. Many journals, books, and theses have been used to gather data, as well as Internet-based information such as Wiley, PubMed, Google Scholar, ScienceDirect, EBSCO, SpringerLink, and Online electronic journals. Key findings Current review elaborates on the potential inhibitory and stimulatory checkpoint pathways which are emerged and have been tested in various preclinical models, clinical trials, and practices. Twenty-odd such significant checkpoints are identified and discussed in the present work. Conclusion A large number of ongoing studies reveal that combination therapies that target more than one signaling pathway may become effective in order to maximize efficacy and minimize toxicity. Moreover, these immunotherapy targets can be a part of integrated therapeutic strategies in addition to classical approaches. It may become a paradigm shift as a promising strategy for cancer treatment.
... Furthermore, the expression levels of CD40 on splenic B cells were significantly increased in nsECT2 and nsECT4 compared to μsECTtreated mice. A few researchers [82,83] have shown that strong CD40 activation promotes B cell differentiation into memory B or plasma cells. However, some studies [84,85] have shown that the differentiation of tumour-associated B cells into regulatory or effector subsets depends on tumour microenvironment. ...
Article
Full-text available
In this work, a time-dependent and time-independent study on bleomycin-based high-frequency nsECT (3.5 kV/cm × 200 pulses) for the elimination of LLC1 tumours in C57BL/6J mice is performed. We show the efficiency of nsECT (200 ns and 700 ns delivered at 1 kHz and 1 MHz) for the elimination of tumours in mice and increase of their survival. The dynamics of the immunomodulatory effects were observed after electrochemotherapy by investigating immune cell populations and antitumour antibodies at different timepoints after the treatment. ECT treatment resulted in an increased percentage of CD4⁺ T, splenic memory B and tumour-associated dendritic cell subsets. Moreover, increased levels of antitumour IgG antibodies after ECT treatment were detected. Based on the time-dependent study results, nsECT treatment upregulated PD 1 expression on splenic CD4⁺ Tr1 cells, increased the expansion of splenic CD8⁺ T, CD4⁺CD8⁺ T, plasma cells and the proportion of tumour-associated pro inflammatory macrophages. The Linˉ population of immune cells that was increased in the spleens and tumour after nsECT was identified. It was shown that nsECT prolonged survival of the treated mice and induced significant changes in the immune system, which shows a promising alliance of nanosecond electrochemotherapy and immunotherapy.
Article
Full-text available
CD19 chimeric antigen receptor T (CD19CAR-T) cells have achieved promising outcomes in relapsed/refractory B-cell malignancies. However, recurrences occur due to the loss of CAR-T cell persistence. We developed dual T/B-cell co-stimulatory molecules (CD28 and CD40) in CAR-T cells to enhance intense tumoricidal activity and persistence. CD19.28.40z CAR-T cells promoted pNF-kB and pRelB downstream signaling while diminishing NFAT signaling upon antigen exposure. CD19.28.40z CAR-T cells demonstrated greater proliferation which translated into effective anti-tumor cytotoxicity in long-term co-culture assay. Repetitive weekly antigen stimulation unveiled continuous CAR-T cell expansion while preserving central memory T-cell subset and lower expression of exhaustion phenotypes. The intrinsic genes underlying CD19.28.40z CAR-T cell responses were compared to conventional CARs and demonstrated the up-regulated genes associated with T-cell proliferation and memory as well as down-regulated genes related to apoptosis, exhaustion, and glycolysis pathway. Enrichment of genes toward T-cell stemness, particularly SELL, IL-7r, TCF7, and KLF2, was observed. Effective and continuing anti-tumor cytotoxicity in vivo was exhibited in both B-ALL and B-NHL xenograft models while demonstrating persistent T-cell memory signatures. The functional enhancement of CD37.28.40z CAR-T cell activities against CD37+ tumor cells was further validated. The modification of dual T/B-cell signaling molecules remarkably maximized the efficacy of CAR-T cell therapy.
Article
The hostile tumor microenvironment limits the efficacy of adoptive cell therapies. Activation of the Fas death receptor initiates apoptosis and disrupting these receptors could be key to increasing CAR T cell efficacy. We screened a library of Fas-TNFR proteins identifying several novel chimeras that not only prevented Fas ligand-mediated kill, but also enhanced CAR T cell efficacy by signaling synergistically with the CAR. Upon binding Fas ligand, Fas-CD40 activated the NF-κB pathway, inducing greatest proliferation and IFN-γ release out of all Fas-TNFRs tested. Fas-CD40 induced profound transcriptional modifications, particularly genes relating to the cell cycle, metabolism, and chemokine signaling. Co-expression of Fas-CD40 with either 4-1BB- or CD28-containing CARs increased in vitro efficacy by augmenting CAR T cell proliferation and cancer target cytotoxicity, and enhanced tumor killing and overall mouse survival in vivo. Functional activity of the Fas-TNFRs were dependent on the co-stimulatory domain within the CAR, highlighting crosstalk between signaling pathways. Furthermore, we show that a major source for Fas-TNFR activation derives from CAR T cells themselves via activation-induced Fas ligand upregulation, highlighting a universal role of Fas-TNFRs in augmenting CAR T cell responses. We have identified Fas-CD40 as the optimal chimera for overcoming Fas ligand-mediated kill and enhancing CAR T cell efficacy.
Article
Full-text available
Engagement of CD40 by its ligand CD154 induces IL-6 production by B lymphocytes. We previously reported that this IL-6 production is dependent upon binding of the adapter protein TNF receptor-associated factor 6 to the cytoplasmic domain of CD40, while binding of TNF receptor-associated factors 2 and 3 is dispensable, as is the activation-induced nuclear translocation of NF-κB. The present study was designed to characterize CD40-mediated transcriptional control of the IL-6 gene in B cells. CD40 engagement on B lymphocytes activated the IL-6 promoter, and mutations in the putative binding sites for AP-1 and C/EBP transcription factors reduced this activation. Interestingly, a mutation in the putative NF-κB binding site completely abrogated the basal promoter activity, thus also rendering the promoter unresponsive to CD40 stimulation, suggesting that this site is required for binding of NF-κB constitutively present in the nucleus of mature B cells. The expression of dominant negative Fos or C/EBPα proteins, which prevent binding of AP-1 or C/EBP complexes to DNA, also reduced CD40-mediated IL-6 gene expression. Furthermore, CD40 stimulation led to phosphorylation of c-Jun on its activation domain, implicating CD40-mediated Jun kinase activation in the transcriptional regulation of IL-6 production.
Article
Full-text available
T cell activation requires a threshold amount of TCR-mediated signals, an amount that is reduced by signals mediated through costimulatory molecules expressed on the T cell surface. Here the role of TNFR2 (p75) as a putative costimulatory receptor for T cell activation was examined. It was found that p75 deficiency in CD8+ T cells increased the requirements for TCR agonist approximately 5-fold. Furthermore, p75−/− T cells display a marked reduction in the proliferative response to TCR agonist. This hypoproliferative response was associated with delayed kinetics of induction of the acute activation markers CD25 and CD69 as well as a marked decrease in the production of IL-2 and IFN-γ. The net result is that very few cells are recruited into the dividing population. Interestingly, CD28 costimulation was only partially effective in rescuing the proliferative defect of p75−/−CD8+ T cells. Thus, p75 provides an important costimulatory signal in addition to that provided by CD28 toward optimal T cell proliferation.
Article
Full-text available
Interferon-γ (IFN-γ) is a pleiotropic lymphokine whose production is restricted to activated T cells and NK cells. Along with other cytokines, IFN-γ gene expression is inhibited by the immunosuppressant cyclosporin A. We have previously identified an intronic enhancer region (C3) of the IFN-γ gene that binds the NF-κB protein c-Rel and that shows partial DNA sequence homology with the cyclosporin A-sensitive NFAT binding site and the 3′-half of the NF-κB consensus site. Sequence analysis of the IFN-γ promoter revealed the presence of two additional C3-related elements (C3-1P and C3-3P). In addition, an NF-κB site (IFN-γ κB) was identified within the promoter region. Based on this observation, we have analyzed the potential role of NF-κB and NFAT family members in regulating IFN-γ transcription. Electrophoretic mobility shift assay analysis demonstrated that after T cell activation, the p50 and p65 NF-κB subunits bind specifically to the newly identified IFN-γ κB and C3-related sites. In addition, we identified the NFAT proteins as a component of the inducible complexes that bind to the C3-3P site. Site-directed mutagenesis and transfection studies demonstrate that calcineurin-inducible transcriptional factors enhance the transcriptional activity of the IFN-γ promoter through the cyclosporin-sensitive C3-3P site, whereas NF-κB proteins functionally interact with the C3-related sites. In addition, when located downstream to the β-galactosidase gene driven by the IFN-γ promoter, the intronic C3 site worked in concert with both the IFN-γ κB and the C3-3P site to enhance gene transcription. These results demonstrate that the coordinate activities of NFAT and NF-κB proteins are involved in the molecular mechanisms controlling IFN-γ gene transcription.
Article
Full-text available
Signaling by Ag to the B cell Ag receptor (BCR) is enhanced by several cooperating signals, including several provided by B-T cell interactions. One of these, CD40, provides critical signals for B cell differentiation, isotype switching, and B cell memory. The molecular mechanisms by which BCR and CD40 signals synergize are not well understood. Although the BCR and CD40 share certain signaling pathways, we hypothesized that unique signals provided by each could provide mutual enhancement of their signaling pathways. The BCR, but not CD40, activates protein kinase D (PKD), while CD40, but not the BCR, employs the TNFR-associated factor (TRAF) adapter proteins in signaling. In this study, we show that genetic or pharmacologic inhibition of BCR-mediated PKD activation in B lymphocytes abrogated the synergy between the CD40 and the BCR, as measured by activation of Ig and cytokine secretion. Interestingly, the role of PKD was dependent upon the association of CD40 with TRAF2, and was inhibited by the binding of TRAF3, revealing a novel functional link between these two classes of signaling molecules.
Article
The nf-kb2 gene encodes the cytoplasmic NF-κB inhibitory protein p100 from which the active p52 NF-κB subunit is derived by proteasome-mediated proteolysis. Ligands which stimulate p100 processing to p52 have not been defined. Here, ligation of CD40 on transfected 293 cells is shown to trigger p52 production by stimulating p100 ubiquitylation and subsequent proteasome-mediated proteolysis. CD40-mediated p52 accumulation is dependent on de novo protein synthesis and triggers p52 translocation into the nucleus to generate active NF-κB dimers. Endogenous CD40 ligation on primary murine splenic B cells also stimulates p100 processing, which results in the delayed nuclear translocation of p52–RelB dimers. In both 293 cells and primary splenic B cells, the ability of CD40 to trigger p100 processing requires functional NF-κB-inducing kinase (NIK). In contrast, NIK activity is not required for CD40 to stimulate the degradation of IκBα in either cell type. The regulation of p100 processing by CD40 is likely to be important for the transcriptional regulation of CD40 target genes in adaptive immune responses.
Article
To address elements that might uniquely characterize CD40 mediated signaling, the nuclear expression of three transcription factors was evaluated following B cell stimulation by CD40L and by anti-lg antibody, Cross-linked CD40L was found to induce nuclear expression of NF-kappa B, AP-1 and NF-AT with a time course and intensity similar to that produced by anti-lg, Examination of NF-kappa B in more detail demonstrated that the CD40 mediated expression of DNA binding complexes correlated with induction of trans-activating activity which again attained similar levels following cross-linking of CD40 and slg, Despite the marked similarity in transcription factor induction triggered through CD40 and slg, differences in the intracellular signaling pathways utilized were apparent in that protein kinase C (PKC) depletion did not affect CD40 mediated induction of NF-kappa B even as induction by anti-lg was abolished, These results suggest that a 'final common pathway' or convergence of transcription factor induction may exist for two distinct receptors, each of which is individually capable of triggering cell cycle progression, despite the use of separate intracellular signaling pathways that differ at the level of PKC. Although transcription factor induction was similar for CD40L and anti-lg early on, subtle differences in expressed NF-kappa B and AP-1 nucleoprotein complexes were apparent at 24 h. Such differences may play a role in determining the variant effects on B cells of stimulation through these two receptors.
Article
The role of Vav in the transcriptional regulation of the human interferon-γ (IFN-γ) promoter was investigated. Overexpression of Vav in Jurkat-TAg cells enhanced T cell receptor (TCR)-induced activation of a luciferase (Luc) reporter gene construct driven by cis-regulatory element of the IFN-γ gene (−346 to +7). Electrophoresis mobility shift and Luc reporter assays demonstrated that the DNA-binding and transcriptional activity of the proximal AP-1-dependent NFAT site (positions −172 to −138), the AP-1/Ying-Yang 1 (YY1)-binding site (−209 to −184), and a consensus AP-1-binding site were upregulated by Vav. Vav enhanced TCR-induced activation of c-Jun N-terminal kinase (JNK) and its upstream regulator, Rho family GTPases. Finally, coexpression of a dominant-negative Rac1 mutant suppressed Vav-mediated upregulation of the transcriptional and DNA-binding activity of the proximal NFAT/AP-1 site and the AP-1/YY1 site, as well as the complete IFN-γ promoter activity. Vav activates the IFN-γ promoter via upregulation of AP-1-binding through a Rac1/JNK pathway.
Article
BCR engagement initiates intracellular calcium ([Ca2+]i) mobilization which is critical for the activation of multiple transcription factors including NF-κB and NFAT. Previously, we showed that Bruton's tyrosine kinase (BTK)-deficient (btk−/−) B cells, which display a modestly reduced calcium response to BCR crosslinking, do not activate NF-κB. Here we show that BTK is also essential for the activation of NFAT following BCR engagement. Pharmacological mobilization of [Ca2+]i in BTK-deficient DT40 B cells (DT40.BTK) does not rescue BCR directed activation of NF-κB and only partially that of NFAT, suggesting existence of additional BTK-signaling pathways in this process. Therefore, we investigated a requirement for BTK in the production of diacylglycerol (DAG). We found that DT40.BTK B cells do not produce DAG in response to BCR engagement. Pharmacological inhibition of PKC isozymes and Ras revealed that the BCR-induced activation of NF-κB requires conventional PKCβ, whereas that of NFAT may involve non-conventional PKCδ and Ras pathways. Consistent with an essential role for BTK in the regulation of NFAT, B cells from btk−/− mice display defective expression of CD5, a gene under the control of NFAT. Together, these results suggest that BCR employs distinct BTK-dependent molecular mechanisms to regulate the activation of NF-κB versus NFAT.
Article
Background. The counter receptors intercellular adhesion molecule (ICAM)-1 and lymphocyte function-associated antigen (LFA)-1 are lymphocyte cell surface adhesion proteins the interaction of which can provide signals for T cell activation. This binding event is important in T cell function, migration, and general immune system regulation. The ability to inhibit this interaction with monoclonal antibodies has proved to be therapeutically useful for several allograft rejection and autoimmune disease models. Methods. Short peptides representing counter-receptor contact domains of LFA-1 and ICAM-1 were examined for their ability to inhibit T cell adhesion and T cell function. Results. Peptides encompassing amino acids Q1-C21 and D26-K50 of ICAM-1, I237-I261 and G441-G466 of the LFA-1 α-subunit, and D134-Q159 of the LFA-1 β-subunit inhibited LFA-1/ICAM-1-dependent adhesion in a phorbol-12,13-dibutyrate-induced model of tonsil T cell homotypic adhesion. This inhibition was specific to the peptide sequence and occurred without stimulation of T cell proliferation. The peptides also were effective in preventing T cell function using a one-way mixed lymphocyte reaction model for bone marrow transplantation. Conclusions. Our data suggest that these peptides or their derivatives may be useful as therapeutic modulators of LFA-1/ICAM-1 interaction during organ transplants.