ArticlePDF Available

Overexpression of Interleukin (IL)-7 Leads to IL-15–independent Generation of Memory Phenotype CD8+ T Cells

Rockefeller University Press
Journal of Experimental Medicine (JEM)
Authors:

Abstract and Figures

Transgenic (TG) mice expressing a high copy number of interleukin (IL)-7 cDNA under the control of the major histocomaptability complex (MHC) class II promoter display a 10-20-fold increase in total T cell numbers. Here, we show that the increase in T cell numbers in IL-7 TG mice is most apparent at the level of memory phenotype CD44hi CD122hi CD8+ cells. Based on studies with T cell receptor (TCR) TG mice crossed to IL-7 TG mice, increased levels of IL-7 may provide costimulation for TCR recognition of self-MHC ligands and thus cause naive CD8+ cells to proliferate and differentiate into memory phenotype cells. In addition, a marked increase in CD44hi CD122hi CD8+ cells was found in IL-7 TG IL-15(-) mice. Since these cell are rare in normal IL-15(-) mice, the dependency of memory phenotype CD8+ cells on IL-15 can be overcome by overexpression of IL-7.
Content may be subject to copyright.
J. Exp. Med.
The Rockefeller University Press • 0022-1007/2002/06/1533/7 $5.00
Volume 195, Number 12, June 17, 2002 1533–1539
http://www.jem.org/cgi/doi/10.1084/jem.20020067
1533
Overexpression of Interleukin (IL)-7 Leads to
IL-15–independent Generation of Memory Phenotype
CD8
T Cells
William C. Kieper,
1
Joyce T. Tan,
1
Brea Bondi-Boyd,
1
Laurent Gapin,
2
Jonathan Sprent,
1
Rhodri Ceredig,
3
and Charles D. Surh
1
1
Department of Immunology, The Scripps Research Institute, La Jolla, CA 92037
2
Division of Developmental Immunology, The La Jolla Institute for Allergy and Immunology,
San Diego, CA 92121
3
U548 INSERM, CEA-Grenoble, F-38054 Grenoble, France
Abstract
Transgenic (TG) mice expressing a high copy number of interleukin (IL)-7 cDNA under the
control of the major histocomaptability complex (MHC) class II promoter display a 10–20-fold
increase in total T cell numbers. Here, we show that the increase in T cell numbers in IL-7 TG
mice is most apparent at the level of memory phenotype CD44
hi
CD122
hi
CD8
cells. Based
on studies with T cell receptor (TCR) TG mice crossed to IL-7 TG mice, increased levels of
IL-7 may provide costimulation for TCR recognition of self-MHC ligands and thus cause naive
CD8
cells to proliferate and differentiate into memory phenotype cells. In addition, a marked
increase in CD44
hi
CD122
hi
CD8
cells was found in IL-7 TG IL-15
mice. Since these cell
are rare in normal IL-15
mice, the dependency of memory phenotype CD8
cells on IL-15
can be overcome by overexpression of IL-7.
Key words: T lymphocytes • homeostasis • cytokines • mice • transgenic
Introduction
The overall size and composition of the mature T cell pool
are closely regulated by homeostatic mechanisms (1). Re-
cently, it has been shown that homeostasis of naive T cells
is primarily controlled through TCR contact with self-
MHC/peptide ligands and exposure to the cytokine IL-7
(2–5). Thus, in the absence of interaction with either of
these ligands, naive T cells disappear and fail to undergo
“homeostatic” proliferation in T cell–deficient conditions
(2–5). Since the density of self-MHC/peptide ligands is
constant, IL-7 may be the key factor in determining the
overall size of the naive T cell pool. Accordingly, overpro-
duction of IL-7 would presumably lead to a proportional
enlargement of the naive T cell pool. In contrast to naive
cells, homeostasis of memory T cells is known to be regu-
lated independently of contact with MHC molecules (6, 7).
In terms of cytokine requirements, none of the common
chain (
c)
*
family of cytokines (IL-2, -4, -7, -9, -15) appears
essential for memory CD4
cells (8). By contrast, memory
CD8
cells are heavily dependent on IL-15 (9, 10). Thus,
most memory CD8
cells express elevated levels of the
IL-15R CD122, undergo selective proliferation in response
to IL-15, and are markedly depleted in IL-15
and IL-15R
mice (9–12).
IL-7 was initially described as a growth factor for B cell
progenitors and was shown to be produced by bone marrow
stromal cells (13). Subsequent studies showed that IL-7 is
also synthesized by other tissues, including thymic and in-
testinal epithelial cells (14, 15). Furthermore, a generation
of IL-7
and IL-7R
mice revealed that IL-7 has a nonre-
dundant role in supporting early development of both B
and T cells (16, 17). In addition to knockout mice, four in-
dependent lines of transgenic (TG) mice overexpressing
IL-7 were generated in the early 1990s (18–21). Although
most of the TG mice were found to possess increased num-
bers of B and T cells, a careful examination of the T cell
pool was not performed.
To study the effect of IL-7 overproduction on T cell ho-
meostasis, we analyzed a B6 TG line that expresses a high
copy number of IL-7 cDNA under the control of the
MHC class II promoter (21). Despite a marked (25–50-
fold) elevation in IL-7 levels, these mice appear healthy
Address correspondence to Charles D. Surh, Dept. of Immunology,
IMM-26, The Scripps Research Institute, 10550 North Torrey Pines Rd.,
La Jolla, CA 92037. Phone: 858-784-2006; Fax: 858-784-8227; E-mail:
csurh@scripps.edu
*
Abbreviations used in this paper:
c, common
chain; LM, littermate;
TG, transgenic.
1534
IL-7 Regulation of Memory CD8
Cell Homeostasis
and, unlike other lines (18, 20), remained free of dermatitis
(21); tumor formation is low and restricted to B cells. Here,
we show that the massive overproduction of T cells in IL-7
TG mice is largely skewed to memory-phenotype CD44
hi
CD122
hi
CD8
cells. Based on the results of crossing IL-7
TG mice to TCR TG and IL-15
mice, IL-7 may play an
important role in guiding both the generation and mainte-
nance of memory CD8
cells.
Materials and Methods
Mice.
IL-7 TG mice (21) were obtained from J. Andersson at
the Basel Institute and maintained by breeding to C57BL/6 (B6)
or B6.Ly 5.1
mice. All mice used in the current study were
hemizygous for the transgene. B6 and B6.PL congenic mice were
purchased from The Scripps Research Institute (TSRI) breeding
facility. Origins of OT-I, HY, IL-15
, and
2m
K
D
mice
were described previously (3, 5).
FACS
®
Analysis.
Suspensions of thymus, LN, and spleen
cells were double- and triple-stained as described previously (2)
using various combinations of PE-anti-CD44, PE-anti-CD122,
Cy5-anti-CD8, FITC-anti-CD8, Cy5-anti-CD4, Cy5-anti-Ly
5.1, and biotinylated T3.70 (all from eBioscience). Biotinylated
anti–Thy-1.1, anti-
c and FITC-V
5, and PE-V
2 antibodies
were purchased from BD PharMingen. Biotinylated mAbs were
detected with Cy5-conjugated streptavidin (Jackson ImmunoRe-
search Laboratories).
For intracellular cytokine staining, spleen cells (4
10
6
cells
per milliliter) in complete medium containing 0.67
ls/ml
Brefeldin A (GolgiStop; BD PharMingen) were incubated for 7 h
in 24-well plates coated with anti-CD3 (2C11; eBioscience)
mAb (0.5
g/ml) and stained as described previously (22).
Briefly, cells were first stained with Cy5-anti-CD8 and PE-anti-
CD44, fixed with paraformaldehyde followed by permeabiliza-
tion with saponin and stained with FITC-anti–IFN-
mAb
(eBioscience).
To assess homeostatic proliferation, small numbers (10
6
cells per
mouse) of FACS
®
-sorted CD44
hi
CD8
cells were labeled with
CFSE (Molecular Probes) and injected into mice exposed to 600
cGy irradiation and analyzed 7 d later as described previously (2).
The rate of background T cell turnover in adult thymecto-
mized mice was performed by adding BrdU (0.8 mg/ml) into the
drinking water for 7 d and staining LN cells as described previ-
ously (23).
Effect of cytokines in vitro was assessed by culturing CD4
and
CD8
cells (2
10
6
cells per milliliter), purified as described pre-
viously (2), for 5 d in complete medium containing IL-7 and/or
IL-15 at 20 ng/ml. Some cultures contained mAbs to IL-7R
(A7R34) (24) and IL-2R
(TM-
1) (25) at 50
g/ml. The in
vivo effect of blocking IL-7R was measured by injecting mice in-
traperitoneally with either 200
g of salt-precipitated A7R34 as-
cites or 200
g of rat IgG every other day over a 7-d period. In
vitro–cultured cells and LN and spleen cells from antibody-
injected mice were analyzed by triple staining with FITC-anti-CD4,
Cy5-anti-CD8 and PE-anti-CD44, or PE-anti-CD122 mAbs.
CDR3 Lengths Analysis of TCR Chains.
Spectratyping of
TCR-V
and -V
repertoire was performed on purified CD4
and CD8
cells as described previously (26). Briefly, after cDNA
synthesis from total RNA, the RNA was denatured and reverse
transcribed. Small aliquots of the cDNA were amplified by PCR
using primers previously described to analyze the repertoire of V
and V
chains (27). The PCR product was then subjected to run-
off reaction with fluorescent-tagged primers and analyzed by cap-
illary electrophoresis in an automated DNA sequencer (Applied
Biosystems) and the distribution of CDR3 lengths determined.
Results and Discussion
Characterization of T Cells in IL-7 TG Mice.
As reported
previously (21), young adult IL-7 TG mice were found to
contain massive numbers of T (and B) cells (10–20 times
above normal) in the secondary lymphoid tissues, despite
the presence of a relatively normal sized thymus (Fig. 1 A).
The increase in peripheral T cell numbers was more pro-
nounced for CD8
cells than CD4
cells and resulted in a
reduction in the CD4/CD8 ratio. The increase in total T
cell numbers and the preferential effect of IL-7 on CD8
cells are likely to be a consequence of IL-7 acting directly
on mature T cells as similar effects were observed in normal
mice given repeated injection of recombinant IL-7 (28–
Figure 1. Characterization of T cells in young adult IL-7 TG mice. (A)
T cell subset distribution and cell numbers in the thymus, LN, and spleen
in IL-7 TG and LMs. Cells were triple-stained for CD4, CD8, CD44, or
CD122 and analyzed by flow cytometry as described in Materials and
Methods. Total numbers of thymocytes and indicated naive and memory
phenotype CD4 and CD8 cells, combined from spleen and pooled
LNs, IL-7 TG, and LMs are shown. Data are representative of 20 TG
and LM mice. (B) Expression of CD44 and CD122 on gated LN CD4
and CD8 cells from IL-7 TG and LMs. (C) Randomly distributed
CDR3 lengths of TCR-V chains expressed on CD8 cells from IL-7
TG mice. CDR3 lengths on most V and all V chains were determined
on purified CD8 cells from IL-7 TG and LM mice as described in Mate-
rials and Methods. Representative data for indicated V chains are shown;
all analyzed chains were randomly distributed in both TG and LM cells.
1535
Kieper et al.
30). Strikingly, the majority (55–70%) of CD8
cells in IL-7
TG mice had a memory (CD44
hi
and CD122
hi
) pheno-
type, whereas most CD4
cells had a naive (CD44
lo
) phe-
notype (Fig. 1 B). For CD8
cells, the increase in total cell
numbers was
50-fold for CD122
hi
cells and
25-fold for
CD44
hi
cells but only
4-fold for naive CD44
lo
cells. For
CD4
cells, CD44
hi
cells were elevated by
5-fold and
CD44
lo
cells by
4-fold (Fig. 1 A). Expression of other
markers, such as CD25, CD62L, CD69, IL-7R
, on TG T
cells was comparable to wild-type T cells (data not shown).
To determine whether the enlargement of the mature T
cell pool is due to an oligoclonal expansion, the TCR rep-
ertoire of purified IL-7 TG CD4
and CD8
cells was ex-
amined. A previous study did not reveal any significant
skewing of the TCR-V
repertoire (21), but it was still
possible that an oligoclonal expansion could be revealed
from analysis of the CDR3 lengths. Examination of all V
chains and the majority of V
chains revealed that the
CDR3 lengths were randomly distributed for all chains
similar to T cells from littermates (LMs) (Fig. 1 C).
The experiments discussed below address the issue of
why IL-7 TG mice show a preferential elevation of mem-
ory phenotype CD8
cells. Why these mice also show a
milder increase in naive phenotype cells, despite having a
normal thymus, will be the subject of another paper.
Functional Capacity of IL-7 TG CD44
hi
CD8
Cells. By
two different parameters, TG CD44hi CD8 cells closely
resembled CD44hi CD8 cells from wild-type mice. First,
as measured by BrdU incorporation in thymectomized
mice, the turnover of TG CD44hi CD8 (and CD44hi
CD4) cells was relatively slow, in fact slower than in LMs;
this finding applied both in adult (8-wk) mice (Fig. 2 A)
and also in neonatal (3-wk) mice (data not shown). Hence
the increase in CD44hi CD8 cells in IL-7 TG mice did
not reflect an increase in their rate of division. However,
total numbers of proliferating T cells in IL-7 TG mice were
7–10-fold higher than in wild-type mice. Second, as mea-
sured by division of CFSE-labeled cells, sorted TG CD44hi
CD8 cells transferred to irradiated B6 hosts behaved simi-
larly to normal B6 CD44hi CD8 cells in their rapid rate of
homeostatic proliferation in these T cell–depleted hosts
(Fig. 2 B). Moreover, like normal CD44hi CD8 cells (Fig.
2 C), but unlike naive CD44lo CD8 cells (6), IL-7 TG
CD44hi CD8 cells underwent efficient homeostatic prolif-
eration in MHC class I (2mDK) mice (Fig. 2 C).
By the above parameters, TG CD44hi CD8 cells closely
resembled normal B6 CD44hi CD8 cells. However, dif-
ferent results were obtained for IFN- synthesis. Thus, af-
ter in vitro stimulation with anti-CD3 mAb, the propor-
tion of IFN- synthesizing cells was much higher for TG
CD44hi CD8 cells (49%) than for normal B6 CD44hi
CD8 cells (9%) (Fig. 3 D). Since IL-7 has costimulatory
activity for T cell stimulation and function (30, 31), this
difference could be a reflection of prior exposure of the TG
cells to high levels of IL-7 in vivo.
Generation of TCR TG CD44hi CD8
Cells in IL-7 TG
Mice. T cells with a memory phenotype usually originate
from naive T cells responding to foreign antigen, but can
also arise from naive T cells undergoing homeostatic prolif-
eration to self-ligands (32). To examine which ligands, self
or foreign, cause production of memory phenotype CD8
cells in IL-7 TG mice, we crossed two TCR TG lines,
OT-I and HY, to IL-7 TG mice.
Figure 2. CD44hi CD8 cells from IL-7 TG mice functionally resem-
ble CD44hi CD8 cells from wild-type mice. (A) Background turnover
over of T cells IL-7 TG mice. Adult thymectomized IL-7 TG and LMs
were given BrdU in the drinking water for 7 d and LN cells were triple
stained for BrdU, CD44, and either CD4 or CD8 as described in Materi-
als and Methods. (B) IL-7 TG CD44hi CD8 cells undergo the same rate
of homeostatic proliferation in T cell–depleted hosts as wild-type CD44hi
CD8 cells. CD44hi CD8 cells from Ly 5.1 IL-7 TG and LM mice pu-
rified by cell sorting were CFSE labeled and injected into irradiated (600
cGy) B6 mice. 7 d after transfer donor cells in host LN were analyzed for
CFSE expression by staining for Ly5.1 and CD8. Shown are gated do-
nor LMs and TG CD8 cells. (C) Homeostatic proliferation of IL-7 TG
CD44hi CD8 cells is MHC independent. Sorted and CFSE-labeled
CD44hi CD8 cells from Ly 5.1 IL-7 TG mice were injected into irra-
diated B6 and MHC class I(2mDK) mice and analyzed 7 d later as
in B. (D) IL-7 TG CD44hi CD8 cells synthesize IFN- upon activation.
Spleen cells from IL-7 TG and LMs were stimulated with anti-CD3 mAb
in vitro for 7 h in the presence of Brefeldin A and stained for CD8,
CD44, and intracellular IFN- as described in Materials and Methods.
Shown are gated CD8 cells. All data are representative of results from
three experiments.
1536 IL-7 Regulation of Memory CD8 Cell Homeostasis
For the OT-I line, nearly all (95%) of the CD8 cells
in this line are clonotype-positive (V2 V5) cells and
have a naive CD44lo phenotype (data not shown). These
naive cells appear to have significant affinity for self-
MHC/peptide ligands because exposure to these ligands
after transfer to T cell–depleted hosts causes strong ho-
meostatic proliferation of the cells and a switch to a mem-
ory phenotype (3). When OT-I mice were crossed to IL-7
TG mice, thus producing OT-I IL-7 TG mice, total num-
bers of CD8 cells in these mice were increased about
threefold relative to OT-I LMs that lacked the IL-7 trans-
gene (Fig. 3). Significantly, numbers of CD44hi CD8
cells were 20-fold higher in OT-I IL-7 TG mice than in
normal OT-I LMs. In addition, in contrast to OT-I LMs,
many of the “CD44lo” cells in OT-I IL-7 TG mice were
actually CD44int rather than CD44lo (Fig. 3). This finding
is of interest because OT-I cells show a similar CD44int/hi
phenotype after homeostatic proliferation in T cell–depleted
hosts (3). Since nearly all (95%) of the CD8 cells in
OT-I IL-7 TG mice were V2 V5, therefore, it
would seem highly likely that the stimulus for expansion
of CD44int/hi cells in these mice was provided by self-
MHC/peptide ligands rather than foreign antigens. Thus,
the implication is that high levels of IL-7 in OT-I IL-7
TG mice act as a costimulator to drive naive OT-I cells to
undergo proliferation and differentiation in response to
self-MHC/peptide ligands. Despite the high penetrance of
the OT-I transgene, however, one caveat is that the IL-7
TG OT-I mice were not in a background (recombination
activating gene [RAG]or TCR-) that precludes ex-
pression of endogenous TCR chains. Therefore, the con-
tribution of foreign antigens in upregulation of CD44 can-
not be totally excluded.
If self-MHC/peptide ligands are driving production of
CD44hi CD8 cells in IL-7 TG mice, the capacity of high
levels of IL-7 to augment T cell reactivity to self-MHC/
peptide ligands might not apply to T cells with “below av-
erage” affinity to self-ligands. To assess this possibility we
examined the HY TCR TG line. As shown previously, the
TCR clonotype-positive cells in this line (detected by
T3.70 mAb) fail to undergo homeostatic proliferation in T
cell–depleted hosts (2), presumably because the TCR affin-
ity of T3.70 cells for self-ligands is quite low. HY T cells
do undergo homeostatic proliferation but this applies only
to the T3.70subset (2). As shown in Fig. 3, HY IL-7 TG
mice showed a marked expansion of CD44hi CD8 cells.
Significantly, almost all of these cells were T3.70; there
was no expansion of T3.70 cells and these cells retained a
naive CD44lo phenotype. Thus, expansion of T3.70but
not T3.70 cells in HY IL-7 TG mice correlated closely
with the capacity of T3.70but not T3.70 cells to un-
dergo homeostatic proliferation in T cell–depleted hosts.
Collectively, these data on OT-I and HY mice suggest
that a significant proportion of the expanded CD44hi
CD8 cells in IL-7 TG mice are derived from precursor
cells stimulated by self-MHC/peptide ligands. This notion
is consistent with the finding that the TCR repertoire of
TG CD8 cells appears very diverse (Fig. 1 C). In addition,
however, a proportion of the expanded CD44hi CD8 cells
may arise through the action of IL-7 on naive CD8 cells
responding to foreign antigens. In this respect, we have ob-
served that IL-7 has marked costimulatory activity for T
cells responding to foreign antigens in vivo. Thus, injection
of male splenic APC into IL-7 TG HY mice led to much
greater expansion of T3.70 cells followed by establish-
ment of a larger pool of T3.70 memory CD44hi CD8
cells than was observed in HY mice lacking IL-7 transgene
(data not shown). Although the relative contribution made
by foreign versus self-antigens in the production of CD44hi
CD8 cells in IL-7 TG mice is unknown it is notable that
the marked increase in CD44hi CD8 cell numbers in these
mice is apparent even in neonatal (1.5 wk) mice (data not
shown). With high IL-7 levels, CD44hi CD8 cells arise
very early in life and accumulate progressively.
IL-15–independent Generation of CD44hi CD8
Cells in
IL-7 TG Mice. As mentioned earlier (see Introduction),
survival of most CD44hi CD8 cells in T cell–sufficient
mice has recently been shown to be heavily dependent on
IL-15 (9–11). IL-15 dependency is particularly pronounced
for CD122hi CD44hi CD8 cells, as these cells are conspic-
Figure 3. Di
ff
erentia
l
e
ff
ect o
f
IL-7 overpro
d
uction on
OT-I and HY TCR TG CD8 cells. OT-I and HY TCR
TG mice were bred to express the IL-7 transgene; LN and
spleen cells from 4-wk-old mice double TG mice were
stained for CD8, V2, and CD44 for OT-I mice and
CD8, T3.70, and CD44 for HY female mice. Shown on
the left side for OT-I cells are the profiles of CD44 on
gated CD8 V2 cells from normal OT-I TG mice (thin
line) and IL-7 TG OT-I mice (thick line); CD44 levels on
polyclonal CD8 cells (broken line) from IL-7 TG mice
are shown for comparison. For HY cells, shown are CD44
levels on gated T3.70 cells from normal HY TG mice
(thin line), IL-7 TG HY mice (thick line) and T3.70cells
from IL-7 TG HY mice (broken line). The total numbers
(pooled from LN and spleen) of recovered CD44lo and
CD44hi populations of CD8 cells from the OT-I and HY
mice are shown on the right side. Similar results were
found when the cells were analyzed in terms of CD122 ex-
pression (not shown). Data are representative of results
from four to six individuals for each type of mice.
1537 Kieper et al.
uously absent in IL-15mice (unpublished data). To deter-
mine whether the generation and/or the maintenance of
CD44hi CD8 cells in IL-7 TG mice is IL-15 dependent,
IL-7 TG mice were crossed with IL-15mice. Surpris-
ingly, comparison of 4-wk-old IL-15IL-7 TG mice with
age-matched control IL-15 IL-7 TG LMs revealed that
IL-15 is dispensable for generation of CD122hi CD44hi
CD8 cells in IL-7 TG mice. Thus, in terms of both total
cellularity and phenotype, including both CD44 and
CD122 expression, the peripheral CD8 cell compartment
was comparable in IL-15 IL-7 TG versus IL-15IL-7 TG
mice (Fig. 4). The proportions of CD122hi CD8 and
CD44hi CD81 cells in cell cycle, as determined by BrdU la-
beling, were also comparable between IL-15 versus IL-
15
IL-7 TG mice (data not shown). It should be men-
tioned that IL-15 deficiency was accompanied by a slight
reduction in the total number of CD8 cells and a minor
increase in the total number of CD4 cells (Fig. 4). The
point to emphasize, however, is that numbers of IL-15 re-
sponsive CD122hi CD8 cells rose from nearly undetect-
able levels in IL-15mice to very high levels in IL-7 TG
IL-15mice (Fig. 4).
Since nonTG IL-15 mice are virtually devoid of
CD122hi CD8 cells (reference 11 and Fig. 4), the above
finding with IL-7 TG IL-15mice indicates that high con-
centrations of IL-7 can compensate for absence of IL-15 in
promoting survival of CD44hi CD122hi CD8 cells. In
support of this idea, a submitogenic dose (20 ng/ml) of ei-
ther IL-7 or IL-15 maintained viability of CD44hi CD8
cells in vitro for 5 d, whereas these cells died in the absence
of cytokines (Fig. 5 A). This finding applied equally to T
cells purified from IL-7 TG or wild-type mice. Whereas
IL-7 was also able to enhance survival of other T cell sub-
sets, the antiapoptotic function of IL-15 was largely re-
stricted to CD44hi CD8 cells, presumably because the IL-
15 receptor, CD122, is expressed at a much higher level on
CD44hi CD8 cells than on other T cell subsets (9).
Figure 4. IL-15 is not required for generation of CD44hi CD8 cells in
IL-7 TG mice. LN cells from young adult IL-7 TG mice bred to an IL-15
background were stained for CD4, CD8, and CD44 or CD122 and ana-
lyzed. Shown are histograms of CD44 and CD122 on gated CD4 and
CD8 cells as compared with the same cells from age-matched control
LMs, IL-7 TG, and IL-15mice. Total numbers of CD4, CD8, and
CD122hi CD8 cells from spleen and LNs from the mice are shown be-
low. Data are representative of results from six to eight individuals for
each type of mice.
Figure 5. Role of IL-7 in supporting survival of CD44hi CD8 cells in
vitro and in vivo in IL-7 TG mice. (A) IL-7 can maintain survival of
CD44hi CD8 cells in vitro. Aliquots of purified IL-7 TG LN T cells
were incubated with the indicated cytokines at (20 ng/ml) either alone or
in the presence of anti–IL-7R mAb (A7R34), anti-CD122 mAb
(TM-1), or both mAbs at 50 g/ml. The cells were harvested 5 d later
and stained for CD4, CD8, and CD44, and the numbers of viable CD44hi
CD8 cells were calculated. (B) Blocking the IL-7R leads to depletion of
CD44hi CD8 cells in IL-7 TG mice. Groups of two to three LMs, IL-7
TG, and IL-15IL-7 TG mice were injected three times with either 200
g of anti–IL-7R mAb (A7R34) or rat IgG at 2-d intervals. 2 d after the
last injection, LN and spleen cells were collected and stained for CD4,
CD8, and CD44. The total numbers of CD44hi CD8 cells recovered
from spleens and pooled LN are shown. Another experiment showed
similar results.
1538 IL-7 Regulation of Memory CD8 Cell Homeostasis
Although IL-7 is presumed to mediate its effect through
the IL-7R, it is conceivable that the high levels of IL-7
available in IL-7 TG mice might act through weak cross-
reactive binding to the IL-15R. This possibility seems un-
likely for two reasons. First, the capacity of IL-7 to rescue
CD44hi CD8 cells from spontaneous death in vitro was
strongly inhibited by anti–IL-7R mAb, and the presence
of anti-CD122 mAb did not diminish the effect of IL-7
(Fig. 5 A). This finding also applied when the concentra-
tion of IL-7 was raised to a high level (data not shown).
Second, injecting either IL-7 TG or IL-15IL-7 TG mice
for 1 wk with anti–IL-7R mAb (200 g every other day)
caused a marked reduction in total numbers of CD44hi
CD8 cells in the mice (Fig. 5 B). A partial reduction in
CD44hi CD8 cells numbers was also observed in control
LM mice (Fig. 5 B). The reduction in cell numbers in these
mice is unlikely to be due to antibody-mediated opsoniza-
tion as depletion of other IL-7Rhi cells, e.g., CD44hi CD4
cells and mature B cells, was minimal (data not shown).
Our observation that the elevation of T cell numbers in
IL-7 TG mice is skewed toward CD8 cells is consistent
with prior evidence on the effects of short-term injection
of IL-7 into normal mice (28–30). Although these latter
data indicated that IL-7 acted largely on peripheral T cells
(rather than thymocytes), the issue of whether IL-7 stimu-
lated naive T cells or memory T cells or both subsets was
not resolved. The data in this paper suggest that, in IL-7
TG mice, elevated levels of IL-7 boost homeostasis of
memory CD8 cells through two different mechanisms.
First, the data on IL-7 TG TCR TG mice suggest that
high levels of IL-7 amplify TCR recognition of self-MHC
ligands by naive CD8 cells. For these cells, TCR recogni-
tion of self-ligands in the presence of background levels of
IL-7 normally induces a covert signal that is sufficient to
keep naive CD8 cells alive, but without inducing prolifer-
ation or a change in surface phenotype. With high levels of
IL-7, however, TCR signaling is enhanced and contact
with self-ligands causes naive CD8 cells to proliferate and
differentiate into memory phenotype cells.
Second, the high levels of IL-7 contribute to survival and
turnover of established memory phenotype CD8 cells. As
discussed earlier, the background turnover and survival of
CD44hi CD8 cells is normally controlled by IL-15; IL-7 is
not important, probably because levels of IL-7 in the pe-
ripheral lymphoid tissues of normal mice are too low to
play a significant role in homeostasis. However, in IL-7 TG
mice one can envisage that IL-7 and IL-15 both contribute
to homeostasis. The presence of large numbers of CD44hi
CD122hi CD8 cells in IL-15
IL-7 TG mice but not nor-
mal IL-15mice is then readily explained. Here, high lev-
els of IL-7 compensate for the lack of IL-15 and homeosta-
sis of CD44hi CD8 cells remains normal. The finding that
the total numbers of CD44hi CD8 cells in IL-15IL-7
TG mice dropped sharply after treatment with anti–IL-7R
mAb is consistent with this idea. However, the finding that
anti–IL-7R mAb treatment had a similar effect on IL-15
IL-7 TG mice is surprising because here we expected the
cells to be rescued by IL-15. It is conceivable, however,
that the massive overproduction of CD122hi CD8 cells in
IL-7 TG mice depletes IL-15, thus causing the cells to be
heavily dependent on IL-7.
In addition to memory phenotype CD8 cells, IL-7 TG
mice clearly contain increased numbers of naive T cells, in-
cluding both CD4 and CD8 cells. Likewise, naive T cell
increase in number after injection of IL-7 (28–30). How
IL-7 leads to expansion of naive T cells is unknown, al-
though increases in cell survival, proliferation, and release
of new T cells from the thymus are all likely possibilities.
Resolving this important issue will have to await further
investigation.
We are grateful to Drs. P. Marrack and J. Andersson for sending
A7R34 hybridomas and the IL-7 TG mice, respectively. We also
thank J. Kuhns, M. Chan, and D. Kim for various supports. This is
publication number 14506-IMM from TSRI.
This work was supported by U.S. Public Health Service grants
AI21487, CA38355, and AI46710 (to J. Sprent) and AI41079,
AI45809, and AG20186 (to C.D. Surh). J.T. Tan and W.C. Kieper
are supported by U.S. Public Health Service Institute National Re-
search Service Award HL07196 and AI07244, respectively. C.D.
Surh is a Scholar of the Leukemia and Lymphoma Society.
Submitted: 15 January 2002
Revised: 4 April 2002
Accepted: 9 May 2002
References
1. Goldrath, A.W., and M.J. Bevan. 1999. Selecting and main-
taining a diverse T-cell repertoire. Nature. 402:255–262.
2. Ernst, B., D.-S. Lee, J.M. Chang, J. Sprent, and C.D. Surh.
1999. The peptide ligands mediating positive selection in the
thymus control T cell survival and homeostatic proliferation
in the periphery. Immunity. 11:173–181.
3. Goldrath, A.W., and M.J. Bevan. 1999. Antagonist ligands
for the TCR drive proliferation of mature CD8 T cells in
lymphopenic hosts. Immunity. 11:183–190.
4. Schluns, K.S., W.C. Kieper, S.C. Jameson, and L. Lefrancois.
2000. Interleukin-7 mediates the homeostasis of naive and
memory CD8 T cells in vivo. Nat. Immunol. 1:426–432.
5. Tan, J.T., E. Dudl, E. LeRoy, R. Murray, J. Sprent, K.I.
Weinberg, and C.D. Surh. 2001. IL-7 is critical for homeo-
static proliferation and survival of naive T cells. Proc. Natl.
Acad. Sci. USA. 98:8732–8737.
6. Murali-Krishna, K., L.L. Lau, S. Sambhara, F. Lemonnier, J.
Altman, and R. Ahmed. 1999. Persistence of memory CD8
T cells in MHC class I-deficient mice. Science. 286:1377–
1381.
7. Swain, S.L., H. Hu, and G. Huston. 1999. Class II-indepen-
dent generation of CD4 memory T cells from effectors. Sci-
ence. 286:1381–1383.
8. Lantz, O., I. Grandjean, P. Matzinger, and J.P. Di Santo.
2000. chain required for naive CD4 T cell survival but
not for antigen proliferation. Nat. Immunol. 1:54–58.
9. Zhang, X., S. Sun, I. Hwang, D.F. Tough, and J. Sprent.
1998. Potent and selective stimulation of memory-phenotype
CD8 T cells in vivo by IL-15. Immunity. 8:591–599.
10. Ku, C.C., M. Murakami, A. Sakamoto, J. Kappler, and P.
Marrack. 2000. Control of homeostasis of CD8 memory T
cells by opposing cytokines. Science. 288:675–678.
1539 Kieper et al.
11. Kennedy, M.K., M. Glaccum, S.N. Brown, E.A. Butz, J.L.
Viney, M. Embers, N. Matsuki, K. Charrier, L. Sedger, C.R.
Willis, et al. 2000. Reversible defects in natural killer and
memory CD8 T cell lineages in interleukin 15-deficient
mice. J. Exp. Med. 191:771–780.
12. Lodolce, J.P., D.L. Boone, S. Chai, R.E. Swain, T. Dassop-
oulos, S. Trettin, and A. Ma. 1998. IL-15 receptor maintains
lymphoid homeostasis by supporting lymphocyte homing
and proliferation. Immunity. 9:669–676.
13. Namen, A.E., S. Lupton, K. Hjerrild, J. Wignall, D.Y. Mo-
chizuki, A. Schmierer, B. Mosley, C.J. March, D. Urdal, and
S. Gillis. 1988. Stimulation of B-cell progenitors by cloned
murine interleukin-7. Nature. 333:571–573.
14. Gutierrez, J.C., and R. Palacios. 1991. Heterogeneity of thy-
mic epithelial cells in promoting T-lymphocyte differentia-
tion in vivo. Proc. Natl. Acad. Sci. USA. 88:642–646.
15. Watanabe, M., Y. Ueno, T. Yajima, Y. Iwao, M. Tsuchiya,
H. Ishikawa, S. Aiso, T. Hibi, and H. Ishii. 1995. Interleukin
7 is produced by human intestinal epithelial cells and regu-
lates the proliferation of intestinal mucosal lymphocytes. J.
Clin. Invest. 95:2945–2953.
16. von Freeden-Jeffry, U., P. Vieira, L.A. Lucian, T. McNeil,
S.E.G. Burdach, and R. Murray. 1995. Lymphopenia in in-
terleukin (IL)-7 gene-deleted mice identifies IL-7 as a nonre-
dundant cytokine. J. Exp. Med. 181:1519–1526.
17. Peschon, J.J., P.J. Morrissey, K.H. Grabstein, F.J. Ramsdell,
E. Maraskovsky, B.C. Gliniak, L.S. Park, S.F. Ziegler, D.E.
Williams, C.B. Ware, et al. 1994. Early lymphocyte expan-
sion is severly impaired in interleukin 7 receptor-deficient
mice. J. Exp. Med. 180:1955–1960.
18. Uehira, M., H. Matsuda, I. Hikita, T. Sakata, H. Fujiwara,
and H. Nishimoto. 1993. The development of dermatitis in-
filtrated by  T cells in IL-7 transgenic mice. Int. Immunol.
5:1619–1627.
19. Samaridis, J., G. Casorati, A. Traunecker, A. Iglesias, J.C.
Gutierrez, U. Muller, and R. Palacios. 1991. Development
of lymphocytes in interleukin 7-transgenic mice. Eur. J. Im-
munol. 21:453–460.
20. Rich, B.E., J. Campos-Torres, R.I. Tepper, R.W. Mo-
readith, and P. Leder. 1993. Cutaneous lymphoproliferation
and lymphomas in interleukin 7 transgenic mice. J. Exp.
Med. 177:305–316.
21. Mertsching, E., C. Burdet, and R. Ceredig. 1995. IL-7 trans-
genic mice: analysis of the role of IL-7 in the differentiation
of thymocytes in vivo and in vitro. Int. Immunol. 7:401–414.
22. Prussin, C. 1997. Cytokine flow cytometry: understanding
cytokine biology at the single-cell level. J. Clin. Immunol. 17:
195–204.
23. Tough, D.F., and J. Sprent. 1994. Turnover of naive- and
memory-phenotype T cells. J. Exp. Med. 179:1127–1135.
24. Sudo, T., S. Nishikawa, N. Ohno, N. Akiyama, M. Tamako-
shi, and H. Yoshida. 1993. Expression and function of the in-
terleukin 7 receptor in murine lymphocytes. Proc. Natl. Acad.
Sci. USA. 90:9125–9129.
25. Tanaka, T., M. Tsudo, H. Karasuyama, F. Kitamura, T.
Kono, M. Hatakeyama, T. Taniguchi, and M. Miyasaka.
1991. A novel monoclonal antibody against murine IL-2 re-
ceptor -chain. Characterization of receptor expression in
normal lymphoid cells and EL-4 cells. J. Immunol. 147:2222–
2228.
26. Gapin, L., J.L. Matsuda, C.D. Surh, and M. Kronenberg.
2001. NKT cells derive from double-positive thymocytes
that are positively selected by CD1d. Nat. Immunol. 2:971–
978.
27. Casanova, J.L., P. Romero, C. Widmann, P. Kourilsky, and
J.L. Maryanski. 1991. T cell receptor genes in a series of class
I major histocompatibility complex-restricted cytotoxic T
lymphocyte clones specific for a Plasmodium berghei nonapep-
tide: implications for T cell allelic exclusion and antigen-spe-
cific repertoire. J. Exp. Med. 174:1371–1383.
28. Morrissey, P.J., P. Conlon, K. Charrier, S. Braddy, A. Alpert,
D. Williams, A.E. Namen, and D. Mochizuki. 1991. Admin-
istration of IL-7 to normal mice stimulates B-lymphopoiesis
and peripheral lymphadenopathy. J. Immunol. 147:561–568.
29. Komschlies, K.L., T.A. Gregorio, M.E. Gruys, T.C. Back,
C.R. Faltynek, and R.H. Wiltrout. 1994. Administration of
recombinant human IL-7 to mice alters the composition of
B-lineage cells and T cell subsets, enhances T cell function,
and induces regression of established metastases. J. Immunol.
152:5776–5784.
30. Geiselhart, L.A., C.A. Humphries, T.A. Gregorio, S. Mou, J.
Subleski, and K.L. Komschlies. 2001. IL-7 administration al-
ters the CD4:CD8 ratio, increases T cell numbers, and in-
creases T cell function in the absence of activation. J. Immu-
nol. 166:3019–3027.
31. Morrissey, P.J., R.G. Goodwin, R.P. Nordan, D. Anderson,
K.H. Grabstein, D. Cosman, J. Sims, S. Lupton, B. Acres,
and S.G. Reed. 1989. Recombinant interleukin 7, pre-B cell
growth factor, has costimulatory activity on purified mature
T cells. J. Exp. Med. 169:707–716.
32. Surh, C.D., and J. Sprent. 2000. Homeostatic T cell prolifer-
ation. How far can T cells be activated to self-ligands? J. Exp.
Med. 192:F9–F14.
... 32,33 It was shown that increasing the IL-7 or IL-15 level caused higher memory T cell numbers and longer memory T cell survival. 34,35 In contrast, low IL-7 and IL-15 levels decreased memory T cell numbers. 36,37 Hence, to support a memory phenotype of the infused cells, IL-7 and IL-15 pathways might be upregulated by overexpressing the IL-7 and IL-15 receptors. ...
Article
Full-text available
Several diseases are caused by the lack of functional proteins, including lysosomal storage diseases or haemophilia A and B. Patients suffering from one of these diseases are treated via enzyme replacement therapies to restore the missing protein. Although this treatment strategy prevents some disease symptoms, enzyme replacement therapies are very expensive and require very frequent infusions, which can cause infusion adverse reactions and massively impair the quality of life of the patients. This review proposes a technology to sustainably produce proteins within the patient to potentially make frequent protein‐infusions redundant. This technology is based on blood circulating immune cells as producers of the needed therapeutic protein. To ensure a stable protein concentration over time the cells are equipped with a system, which induces cell proliferation when low therapeutic protein levels are detected and a system inhibiting cell proliferation when high therapeutic protein levels are detected.
... However, tTEM with an IL-7R − CD62L − KLRG1 + phenotype were strongly reduced [36], suggesting that IL-15 derived from Tregs was required for their survival [1] (Figure 1). The fact that the numbers of central memory (T CM ) and effector memory (T EM ) subsets were poorly affected might be due to IL-7R expression on these memory subsets, since IL-7 may compensate for a loss of IL-15 and promote the survival of CD8 + memory T-cells [37,38]. Consistent with the view that IL-15 is particular important for IL-7R − T tEM , this CD8 + memory T-cell subset is strongly reduced in LCMV-infected mice that lack IL-15 in all cells [36]. ...
Article
Full-text available
It is well known that regulatory T‐cells (Tregs) are required to prevent autoimmunity, but they may also have some less‐well understood immune‐stimulatory effects. In particular, in CD8+ T‐cell responses Tregs select high affinity clones upon priming and promote memory by inhibiting inflammation‐dependent generation of short‐lived effector cells. In the current issue of the European Journal of Immunology [Eur. J. Immunol. 2023. 53: XXXX‐XXXX], Madi et al. report the surprising finding that human and murine FOXP3+ Tregs are a physiological relevant source of Interleukin‐15, a homeostatic cytokine that promotes antigen‐independent maintenance of CD8+ memory T‐cells. In mice that lack IL‐15 selectively in FOXP3+Tregs the authors show that the composition of the CD8+ T‐cell memory pool is altered in the absence of Treg‐derived IL‐15, since a subset of terminally effector memory cells is drastically reduced. Otherwise Treg‐derived IL‐15 is dispensable for anti‐viral immune responses and the generation of anti‐viral CD8+ memory T‐cells. These findings add to our understanding of the multifaceted role of Tregs in immune responses, and how IL‐15 derived from different cellular sources maintains anti‐viral T‐cell memory. This article is protected by copyright. All rights reserved
... Therefore, IL7R transcript were mainly contributed by T cells in bone marrows. Several studies demonstrated that both naïve and memory T cells express high levels of IL7R, and IL7 is required for their homoeostasis (32)(33)(34)(35)(36). IL7R was absent on selecting thymocytes and effector cells (37,38). ...
Article
Full-text available
Background Acute myeloid leukemia (AML) with t(8;21) needs to be further stratified. In addition to leukemia cells, immune cells in tumor microenvironment participate in tumor initiation, growth and progression. Interleukins (ILs)/interleukin receptors (ILRs) interaction plays important roles in the antitumor immune response. IL7R is reported to be relevant to prognosis in solid tumor and acute lymphoblastic leukemia. However, the prognostic significance of IL7R in t(8;21) AML remains to be clarified. Methods Bone marrows collected from 156 newly diagnosed t(8;21) AML patients were used for testing IL7R transcript level by TaqMan-based real-time quantitative PCR (RQ-PCR), and RNAseq were performed in 15 of them. Moreover, IL7R expression at diagnosis were measured by RQ-PCR and flow cytometry (FCM) simultaneously in other 13 t(8;21) AML patients. Results t(8;21) AML patients had varied IL7R transcript levels and were categorized into low-expression (IL7R-L) and high-expression (IL7R-H) groups; IL7R-L was significantly associated with a lower relapse-free survival (RFS) rate (P=0.0027) and KITD816/D820 mutation (P=0.0010). Furthermore, IL7R-L was associated with a lower RFS rate in KITD816/D820 group (P=0.013) and IL7R-H/KITD816/D820 patients had similar RFS to KITN822/e8/WT patients (P=0.35). GO analysis enrichment showed that down-regulated genes were predominantly involved in the regulation of T cell and leukocyte activation, proliferation and differentiation in IL7R-L group. IL7R-L had significantly lower levels of Granzymes A/B, CCR7, CD28 and CD27 than IL7R-H group (all P<0.05). FCM analysis showed IL7R protein was primarily expressed in CD4⁺ T and CD8⁺ T cell subset. A significant association was found between the transcript level of IL7R and the percentage of CD8⁺ T cells in nucleated cells (P=0.015) but not CD4⁺ T cells (P=0.47). Conclusion Low IL7R transcript level of bone marrow at diagnosis predicted relapse in t(8;21) AML, which might be caused by the difference in the amount, status and function of T cells.
Article
Full-text available
T cells are an important component of adaptive immunity and protect the host from infectious diseases and cancers. However, uncontrolled T cell immunity may cause autoimmune disorders. In both situations, antigen-specific T cells undergo clonal expansion upon the engagement and activation of antigens. Cellular metabolism is reprogrammed to meet the increase in bioenergetic and biosynthetic demands associated with effector T cell expansion. Metabolites not only serve as building blocks or energy sources to fuel cell growth and expansion but also regulate a broad spectrum of cellular signals that instruct the differentiation of multiple T cell subsets. The realm of immunometabolism research is undergoing swift advancements. Encapsulating all the recent progress within this concise review in not possible. Instead, our objective is to provide a succinct introduction to this swiftly progressing research, concentrating on the metabolic intricacies of three pivotal nutrient classes—lipids, glucose, and amino acids—in T cells. We shed light on recent investigations elucidating the roles of these three groups of metabolites in mediating the metabolic and immune functions of T cells. Moreover, we delve into the prospect of “editing” metabolic pathways within T cells using pharmacological or genetic approaches, with the aim of synergizing this approach with existing immunotherapies and enhancing the efficacy of antitumor and antiinfection immune responses.
Article
Full-text available
Mucosal‐associated invariant T (MAIT) cells have a semi‐invariant T‐cell receptor that allows recognition of antigen in the context of the MHC class I‐related (MR1) protein. Metabolic intermediates of the riboflavin synthesis pathway have been identified as MR1‐restricted antigens with agonist properties. As riboflavin synthesis occurs in many bacterial species, but not human cells, it has been proposed that the main purpose of MAIT cells is antibacterial surveillance and protection. The majority of human MAIT cells secrete interferon‐gamma (IFNg) upon activation, while some MAIT cells in tissues can also express IL‐17. Given that MAIT cells are present in human barrier tissues colonized by a microbiome, MAIT cells must somehow be able to distinguish colonization from infection to ensure effector functions are only elicited when necessary. Importantly, MAIT cells have additional functional properties, including the potential to contribute to restoring tissue homeostasis by expression of CTLA‐4 and secretion of the cytokine IL‐22. A recent study provided compelling data indicating that the range of human MAIT cell functional properties is explained by plasticity rather than distinct lineages. This further underscores the necessity to better understand how different signals regulate MAIT cell function. In this review, we highlight what is known in regards to activating and inhibitory signals for MAIT cells with a specific focus on signals relevant to healthy and inflamed tissues. We consider the quantity, quality, and the temporal order of these signals on MAIT cell function and discuss the current limitations of computational tools to extrapolate which signals are received by MAIT cells in human tissues. Using lessons learned from conventional CD8 T cells, we also discuss how TCR signals may integrate with cytokine signals in MAIT cells to elicit distinct functional states.
Article
OSE-127 is a humanized mAb targeting the IL-7Rα-chain (CD127), under development for inflammatory and autoimmune disease treatment. It is a strict antagonist of the IL-7R pathway, is not internalized by target cells, and is noncytotoxic. In this work, a first-in-human, phase I, randomized, double-blind, placebo-controlled, single-center study was carried out to determine the safety, pharmacokinetics, pharmacodynamics, and immunogenicity of OSE-127 administration. Sixty-three healthy subjects were randomly assigned to nine groups: six single ascending dose groups with i.v. administration (0.002-10 mg/kg), a single s.c. treatment group (1 mg/kg), and two double i.v. injection groups (6 or 10 mg/kg). Subjects were followed during <146 d. OSE-127's pharmacokinetic half-life after a single dose increased from 4.6 (1 mg/kg) to 11.7 d (10 mg/kg) and, after a second dose, from 12.5 (6 mg/kg) to 16.25 d (10 mg/kg). Receptor occupancy was ≥95% at doses ≥0.02 mg/kg, and this saturation level was maintained >100 d after two i.v. infusions at 10 mg/kg. IL-7 consumption was inhibited by OSE-127 administration, as demonstrated by a decreased IL-7 pathway gene signature in peripheral blood cells and by ex vivo T lymphocyte restimulation experiments. OSE-127 was well tolerated, with no evidence of cytokine-release syndrome and no significant alteration of blood lymphocyte counts or subset populations. Altogether, the observed lack of significant lymphopenia or serious adverse events, concomitant with the dose-dependent inhibition of IL-7 consumption by target cells, highlights that OSE-127 may show clinical activity in IL-7R pathway-involved diseases.
Article
Overview IL-7 is a member of the family of cytokines with four anti-parallel α helixes that bind Type I cytokine receptors. It is produced by stromal cells and is required for development and homeostatic survival of lymphoid cells. Genomic architecture Interleukin 7 (IL7) human IL7: gene ID: 3574 on ch 8; murine Il7 gene ID: 16,196 on ch 3. Protein Precursor contains a signal sequence, mature human IL-7 peptide 152aa, predicted 17.4kd peptide, glycosylated resulting in 25kd. Crystal structure: http://www.rcsb.org/structure/3DI2. Regulation of IL-7 production Major producers are stromal cells in thymus, bone marrow and lymphoid organs but also reported in other tissues. Production is primarily constitutive but reported to be affected by IFNγ and other factors. IL-7 receptors Two chains IL-7Rα (IL-7R) and γc (IL-2RG). Human IL-7R: gene ID 3575 on ch 5; human IL2RG: gene ID 3561 on ch X; mouse IL-7R: gene ID 16,197 on ch 15; murine Il2rg gene ID 16,186 on ch X. Member of γc family of receptors for cytokines IL-2, −4, −9, −15, and −21. Primarily expressed on lymphocytes but reports of other cell types. Expression in T-cells downregulated by IL-7. Low expression on Tregs, no expression on mature B-cells. Crystal structure: http://www.rcsb.org/structure/3DI2. IL-7 receptor signal transduction pathways Major signals through JAK1, JAK3 to STAT5 and through non-canonical STAT3, STAT1, PI3K/AKT and MEK/ERK pathways. Biological activity of IL-7 Required for survival of immature thymocytes, naïve T-cells, memory T-cells, pro-B-cells and innate lymphocytes. Pharmacological treatment with IL-7 induces expansion of naïve and memory T-cells and pro-B-cells. Abnormalities of the IL-7 pathway in disease Deficiencies in the IL-7 pathway in humans and mice result in severe combined immunodeficiency due to lymphopenia. Excessive signaling of the pathway in mice drives autoimmune diseases and in humans is associated with autoimmune syndromes including multiple sclerosis, type 1 diabetes, rheumatoid arthritis, sarcoidosis, atopic dermatitis and asthma. Mutations in the IL-7 receptor pathway drive acute lymphoblastic leukemia. Clinical applications IL-7 has been evaluated in patients with cancer and shown to expand lymphocytes. It accelerated lymphocyte recovery after hematopoietic stem cell transfer, and increased lymphocyte counts in AIDS patients and sepsis patients. Monoclonal antibodies blocking the IL-7 receptor are being evaluated in autoimmune diseases. Cytotoxic monoclonals are being evaluated in acute lymphoblastic leukemia. Drugs blocking the signal transduction pathway are being tested in autoimmunity and acute lymphoblastic leukemia.
Article
Full-text available
To investigate the role of interleukin 7 (IL-7) in the development of the lymphoid system, we have generated two lines of transgenic mice carrying an Ib7 eDNA fused to an immunoglobulin heavy chain promoter and enhancer. This transgene is expressed in the bone marrow, lymph nodes, spleen, thymus, and skin provoking a perturbation of T cell development characterized by a marked reduction of CD4+CD8 + (double-positive) thymocytes. Quite unexpectedly, however, both lines also develop a progressive cutaneous disorder involving a dermal lymphoid infiltrate that results in progressive alopecia, hyperkeratosis, and exfoliation. Although the infiltrate is primarily composed of T lineage cells, its development is not impeded in the athymic nu/nu background. Furthermore, the phenotype can be transmitted horizontally by transplanting lymphoid tissues or skin to syngeneic wild-type mice. Thus, the phenotype is conveyed by skin-homing, mobile cells (presumably the infiltrating lymphocytes) in a cell-autonomous fashion. In addition to the skin phenotype, this transgene also provokes the development of a lymphoproliferative disorder that induces B and T cell lymphomas within the first 4 mo of life. These findings suggest potential physiologic actions of IL-7 in T cell development and in cutaneous immunity. They also demonstrate that IL-7 can act as an oncogene in the living organism.
Article
Full-text available
C57BL/6 mice genetically deficient in interleukin 15 (IL-15−/− mice) were generated by gene targeting. IL-15−/− mice displayed marked reductions in numbers of thymic and peripheral natural killer (NK) T cells, memory phenotype CD8+ T cells, and distinct subpopulations of intestinal intraepithelial lymphocytes (IELs). The reduction but not absence of these populations in IL-15−/− mice likely reflects an important role for IL-15 for expansion and/or survival of these cells. IL-15−/− mice lacked NK cells, as assessed by both immunophenotyping and functional criteria, indicating an obligate role for IL-15 in the development and functional maturation of NK cells. Specific defects associated with IL-15 deficiency were reversed by in vivo administration of exogenous IL-15. Despite their immunological defects, IL-15−/− mice remained healthy when maintained under specific pathogen-free conditions. However, IL-15−/− mice are likely to have compromised host defense responses to various pathogens, as they were unable to mount a protective response to challenge with vaccinia virus. These data reveal critical roles for IL-15 in the development of specific lymphoid lineages. Moreover, the ability to rescue lymphoid defects in IL-15−/− mice by IL-15 administration represents a powerful means by which to further elucidate the biological roles of this cytokine.
Article
Full-text available
We report here the first extensive study of a T cell repertoire for a class I major histocompatibility complex (MHC)-restricted cytotoxic T lymphocyte (CTL) response. We have found that the T cell receptors (TCRs) carried by 28 H-2Kd-restricted CTL clones specific for a single Plasmodium berghei circumsporozoite nonapeptide are highly diverse in terms of V alpha, J alpha, and J beta segments and aminoacid composition of the junctional regions. However, despite this extensive diversity, a high proportion of the TCRs contain the same V beta segment. These results are in contrast to most previously reported T cell responses towards class II MHC-peptide complexes, where the TCR repertoires appeared to be much more limited. In our study, the finding of a dominant V beta in the midst of otherwise highly diverse TCRs suggests the importance of the V beta segment in shaping the T cell repertoire specific for a given MHC-peptide complex. As an additional finding, we observed that nearly all clones have rearranged both TCR alpha loci. Moreover, as many as one-third of the CTL clones that we analyzed apparently display two productive alpha rearrangements. This argues against a regulated model of sequential recombination at the alpha locus and consequently raises the question of whether allelic exclusion of the TCR alpha chain is achieved at all.
Article
We report here the first extensive study of a T cell repertoire for a class I major histocompatibility complex (MHC)-restricted cytotoxic T lymphocyte (CTL) response. We have found that the T cell receptors (TCRs) carried by 28 H-2Kd-restricted CTL clones specific for a single Plasmodium berghei circumsporozoite nonapeptide are highly diverse in terms of V alpha, J alpha, and J beta segments and aminoacid composition of the junctional regions. However, despite this extensive diversity, a high proportion of the TCRs contain the same V beta segment. These results are in contrast to most previously reported T cell responses towards class II MHC-peptide complexes, where the TCR repertoires appeared to be much more limited. In our study, the finding of a dominant V beta in the midst of otherwise highly diverse TCRs suggests the importance of the V beta segment in shaping the T cell repertoire specific for a given MHC-peptide complex. As an additional finding, we observed that nearly all clones have rearranged both TCR alpha loci. Moreover, as many as one-third of the CTL clones that we analyzed apparently display two productive alpha rearrangements. This argues against a regulated model of sequential recombination at the alpha locus and consequently raises the question of whether allelic exclusion of the TCR alpha chain is achieved at all.
Article
Interleukin (IL)-7 is a potent stimulus for immature T and B cells and, to a lesser extent, mature T cells. We have inactivated the IL-7 gene in the mouse germline by using gene-targeting techniques to further understand the biology of IL-7. Mutant mice were highly lymphopenic in the peripheral blood and lymphoid organs. Bone marrow B lymphopoiesis was blocked at the transition from pro-B to pre-B cells. Thymic cellularity was reduced 20-fold, but retained normal distribution of CD4 and CD8. Splenic T cellularity was reduced 10-fold. Splenic B cells, also reduced in number, showed an abnormal population of immature B cells in adult animals. The remaining splenic populations of lymphocytes showed normal responsiveness to mitogenic stimuli. These data show that proper T and B cell development is dependent on IL-7. The IL-7-deficient mice are the first example of single cytokine-deficient mice that exhibit severe lymphoid abnormalities.
Article
Transgenic mice constitutively expressing IL-7 developed severe dermatitis with erythroderma and alopecia. The skin lesions were characterized by massive infiltration of mononuclear cells. Immunofluorescence staining showed that most of the infiltrating cells were T cells with the majority bearing the γδ TCR other than the Vγ5 moiety. Furthermore, the number of γδ T cells had increased in the lymphold organs of the dermatitis animals. These findings idicate the strong relationship between the expression of IL-7 and the development of γδ T cells in vivo and the pathological involvement of proliferated and/or activated γδ T cells in skin disease.
Article
We have generated a high copy number transgenic mouse line in which expression of mouse IL-7 cDNA is under the control of the mouse MHC class II E(alpha) promoter, These mice were generated in order to see if IL-7 over-production in the thymus altered either thymocyte differentiation or the process of negative selection, Using in site hybridization, IL-7 transcripts could be detected in the thymic cortex and medulla as well as the spleen and lymph nodes of transgenic mice but was undetectable in normal controls, Phenotypic and molecular analysis of thymocytes from embryonic and adult transgenic mice failed to reveal a dramatic effect of IL-7 on thymocyte differentiation and negative selection of the TCR V-beta repertoire appeared to be intact, In peripheral lymph nodes, there was a massive (30-fold) increase in the number of T cells (CD8(+) > CD4(+)) and simultaneous presence of immature (B220(+), Ig(-)) B cells, TCR repertoire analysis showed that the expansion of peripheral T cells was polyclonal. Using the polymerase chain reaction (PCR), transgene-specific IL-7 transcripts could be detected in the thymus from day 14 of fetal development, However, using semi-quantitative PCR, there was no dramatic increase in the degree of TCR beta or TCR alpha gene rearrangements during thymocyte ontogeny in vivo, Similarly, when fetal mouse thymus lobes were cultured with IL-7 in vitro, there was no dramatic increase in the degree of TCR beta or TCR alpha gene rearrangements, We conclude that IL-7 is probably not an important differentiation factor for immature mouse thymocytes.
Article
Normal mice were injected with IL-7 (500 ng, twice daily) for various periods of time up to 6 days and the cellularity and phenotypic composition of the thymus, spleen, lymph node, and bone marrow was assessed. After 6 days of treatment, significant increases in the cellularity of the spleen, lymph node, and bone marrow were observed which returned to the normal range within 6 days after cessation of treatment. After 3 days of IL-7 treatment, increased numbers of B220+/surface(s) IgM- bone marrow cells were observed. After 6 days of treatment, these numbers were still further increased and a significant population of B220+/sIgM- cells were observed in the spleen. The numbers of c mu+/sIgM- cells were also increased in the IL-7-treated mice. Analysis of the expression of B220 and BP-1 on the sIgM- bone marrow cells revealed that the B220+/BP-1+ population was dramatically increased after IL-7 treatment and the size of the B220+/BP-1- population did not differ from control mice. The pre-B cell numbers declined rapidly after the cessation of IL-7 treatment. After 6 days of IL-7 treatment, a twofold increase in the number of B cells in the spleen and lymph node was observed. The B cell numbers declined to normal values within 6 days after the cessation of IL-7 administration. In the spleens of the IL-7-treated mice, there was a significant increase in the number of B cells with an immature phenotype (e.g., sIgMhi/sIgDlo, decreased levels of Ia and FcR expression). The numbers of CD8+ and CD4+ T cells were also increased in the lymph node and spleen of the IL-7-treated mice. These numbers declined to normal levels after the cessation of IL-7 treatment.
Article
A mAb specific for the murine IL-2R beta-chain (IL-2R beta) was produced by immunizing a rat with a rat transfectant cell line expressing a large number of cDNA-encoded murine IL-2R beta. The mAb, designated TM-beta 1, is specifically reactive with the murine IL-2R beta cDNA-transfectant but not with the recipient cell, and immunoprecipitates murine IL-2R beta of Mr 75 to 85 kDa. TM-beta 1 mAb completely abolished the high affinity IL-2 binding by inhibiting the ligand binding to IL-2R beta. Murine IL-2R beta was found to be constitutively expressed on a subpopulation of CD8+ T cells and almost all NK1.1+ NK cells in the spleen, whereas TM-beta 1 mAb inhibited the proliferation of spleen cells induced by 1 nM of IL-2. Interestingly, EL-4 cells that express murine IL-2R beta as detected by TM-beta 1 mAb can bind neither human nor murine IL-2 under the intermediate affinity conditions, although cDNA-directed human IL-2R beta expressed in the same EL-4 cells has been previously shown to manifest the intermediate affinity IL-2 binding. These results may imply that functional expression of IL-2R beta is differentially regulated between humans and mice. Finally, our neutralizing anti-IL-2R beta mAb TM-beta 1 will be useful not only for various in vitro studies but also for in vivo studies to directly investigate the possible involvement of the IL-2/IL-2R pathway in the generation and differentiation of T lymphocytes and NK cells.