ArticlePDF Available

Human breast cancer cells enhance self tolerance by promoting evasion from NK cell antitumor immunity

Authors:

Abstract and Figures

NK cells are a major component of the antitumor immune response and are involved in controlling tumor progression and metastases in animal models. Here, we show that dysfunction of these cells accompanies human breast tumor progression. We characterized human peripheral blood NK (p-NK) cells and malignant mammary tumor-infiltrating NK (Ti-NK) cells from patients with noninvasive and invasive breast cancers. NK cells isolated from the peripheral blood of healthy donors and normal breast tissue were used as controls. With disease progression, we found that expression of activating NK cell receptors (such as NKp30, NKG2D, DNAM-1, and CD16) decreased while expression of inhibitory receptors (such as NKG2A) increased and that this correlated with decreased NK cell function, most notably cytotoxicity. Importantly, Ti-NK cells had more pronounced impairment of their cytotoxic potential than p-NK cells. We also identified several stroma-derived factors, including TGF-β1, involved in tumor-induced reduction of normal NK cell function. Our data therefore show that breast tumor progression involves NK cell dysfunction and that breast tumors model their environment to evade NK cell antitumor immunity. This highlights the importance of developing future therapies able to restore NK cell cytotoxicity to limit/prevent tumor escape from antitumor immunity.
Content may be subject to copyright.
Research article
The Journal of Clinical Investigation      http://www.jci.org      Volume 121      Number 9      September 2011  3609
Human breast cancer cells enhance
self tolerance by promoting evasion
from NK cell antitumor immunity
Emilie Mamessier,1 Aude Sylvain,1 Marie-Laure Thibult,1 Gilles Houvenaeghel,2
Jocelyne Jacquemier,2 Rémy Castellano,1 Anthony Gonçalves,2 Pascale André,3
François Romagné,3 Gilles Thibault,4 Patrice Viens,2 Daniel Birnbaum,1,2
François Bertucci,1,2 Alessandro Moretta,5 and Daniel Olive1,2
1Centre de Recherche en Cancérologie de Marseille, INSERM UMR U891, Marseille, France. 2Institut Paoli-Calmettes, Marseille, France.
3Innate Pharma, Route de Luminy, Marseille, France. 4UMR CNRS 6239 GICC Immuno-Pharmaco-Génétique des Anticorps thérapeutiques,
Université François Rabelais de Tours, Tours, France. 5Dipartimento di Medicina Sperimentale (DIMES), University di Genova, Genoa, Italy.
NK cells are a major component of the antitumor immune response and are involved in controlling tumor
progression and metastases in animal models. Here, we show that dysfunction of these cells accompanies
human breast tumor progression. We characterized human peripheral blood NK (p-NK) cells and malignant
mammary tumor-infiltrating NK (Ti-NK) cells from patients with noninvasive and invasive breast cancers.
NK cells isolated from the peripheral blood of healthy donors and normal breast tissue were used as controls.
With disease progression, we found that expression of activating NK cell receptors (such as NKp30, NKG2D,
DNAM-1, and CD16) decreased while expression of inhibitory receptors (such as NKG2A) increased and that
this correlated with decreased NK cell function, most notably cytotoxicity. Importantly, Ti-NK cells had more
pronounced impairment of their cytotoxic potential than p-NK cells. We also identified several stroma-derived
factors, including TGF-β1, involved in tumor-induced reduction of normal NK cell function. Our data there-
fore show that breast tumor progression involves NK cell dysfunction and that breast tumors model their envi-
ronment to evade NK cell antitumor immunity. This highlights the importance of developing future therapies
able to restore NK cell cytotoxicity to limit/prevent tumor escape from antitumor immunity.
Introduction
Breast cancer (BC) is the primary cause of cancer deaths in women. 
The main cause of this mortality is the metastatic spread to other 
organs (1). Metastasis occurs when  tumor cells acquire invasive 
features (2) and the ability to escape from antitumor immunity (3, 
4). Defects in antitumor immunity may also facilitate BC occur-
rence. Indeed, mice deficient in IFN-γ production spontaneously 
develop mammary tumors (5). Breast tumor cells transplanted into 
NOD/SCID mice (which lack adaptive immunity) form noninvasive 
tumors, whereas the same cells transplanted into NOD/SCID/γ-cnull
mice (no adaptive immunity and no NK cells) form invasive tumors 
that metastasize rapidly (6). This effect is strictly dependent on NK 
cells (7). Similarly, in a highly metastatic model, BC metastasized to 
the lung only after elimination of NK cells by Tregs (8).
Advanced BC patients show defects in antitumor immunity, such 
as alterations of DC maturation (9) and an increase in Treg infil-
trates (10). Major impairment of peripheral blood NK cell matura-
tion and cytotoxic functions has also been reported in metastatic 
BC (11). Several gene expression profiling studies have shown that 
a better outcome is associated with a strong cytotoxic infiltrate 
containing NK cells (12–15). These data suggest that BC progres-
sion is linked to antitumor immunity efficiency and particularly 
to NK cells. However, the precise relationships between NK cells 
and BC progression in humans have not been studied so far.
NK cells are innate immune cells that have the natural abil-
ity to  distinguish normal cells  from “modified” cancer cells 
(16).  Once  activated,  NK cells eliminate their targetthrough 
the release of cytotoxic enzymes (perforin 1, granzymes, granu-
lysin) and/or soluble factors (chemokines and inflammatory 
cytokines), which, in turn, recruit and/or activate other effectors 
(17). Activating and inhibitory receptors present on NK cells are 
triggered during target cell recognition and induce a positive 
or a negative cell signaling pathway, respectively. The integra-
tion  ofthese  opposite  signals determines NKcell activation 
(18). The main activating receptors or coreceptors of NK cells 
are NKG2D, the natural cytotoxicity receptors (NCRs) NKp30 
and NKp46, DNAM-1, CD2, NKp80, 2B4, and NTBA (19–21). 
These molecules recognize various ligands usually upregulated 
upon cellular stress (22). NK cells also express the Fc immuno-
globulin fragment low-affinity receptor or CD16, which, when 
cross-linked, induces a powerful reaction called antibody-depen-
dent cellular cytotoxicity (ADCC). Inhibitory receptors include 
the killer immunoglobulin receptors (KIRs), NKG2A, CD85j, 
and LAIRs (23–26). They are specific for different HLA–class I 
molecules. Accordingly, NK cells can kill target cells that have 
lost (or express low amounts of) HLA–class I molecules, which 
is frequently the case for tumor cells, including breast tumor 
cells (27). However, tumor cells also have the ability to impair 
NK cell “visibility” through the modulation of their receptors 
(28, 29). Recent studies have shown that several molecules, nota-
bly inhibitory factors often found in the tumor microenviron-
ment, such as IDO1 and TGF-β1, can sharply impair NK cells’ 
phenotype and functions (30, 31).
Conflict of interest: Pascale André and François Romagné are full-time employees at 
Innate Pharma, a biopharmaceutical company, and Alessandro Moretta is a founder 
of and a shareholder in Innate Pharma.
Citation for this article:J Clin Invest. 2011;121(9):3609–3622. doi:10.1172/JCI45816.
research article
3610 The Journal of Clinical Investigation      http://www.jci.org      Volume 121      Number 9      September 2011
We show here that mechanisms of escape from NK cell–mediat-
ed immunity are at play in BC patients. In a cohort of BC patients 
sampled at different stages of the malignant process, we found 
that breast tumors have altered NK cell phenotype and function 
and that invasive tumors build strong inhibitory microenviron-
ments to escape NK cell antitumor immunity.
Results
p-NK cell phenotype is altered in invasive BC patients
We prospectively enrolled patients with different stages of BC at 
diagnosis. Based on tumor pathologic tumor-node-metastasis 
(pTNM) classification and tumor margins, groups were composed 
of noninvasive (in situ) BCs (Tis) (n = 8) and invasive BCs (n = 113) 
including localized (LOC) (n = 55), locally advanced (LA) (n = 26), 
and metastatic stages (M) (n = 32). This classification, proposed by 
the American Joint Committee on Cancer (AJCC), is based on the 
histoclinical extension of disease, which represents the most impor-
tant prognostic factor of BC, with decreasing survival from the Tis 
group to the M group (32). Clinical characteristics of the patients 
are summarized in Table 1. Due to the prospective nature of our 
study, the clinical follow-up of the patients was too short for sur-
vival analysis. Patients with benign mammary tumors (B) (n = 19) 
and healthy donors (HD) (n = 22) were included as control groups.
To explore NK cell cytotoxic potential, we first determined the 
expression of 22 NK cell receptors on cells from fresh blood sam-
ples. The MFI of NKp30, NKG2D, DNAM-1, and 2B4 were lower 
in patients with invasive breast tumors (LOC, LA, and M) than in 
control groups and patients with Tis tumors (Figure 1, A–D). The 
expression of CD16 was lower in patients with invasive BC than in 
control groups (Figure 1E). The expression of other activating mol-
ecules did not vary with tumor progression (Figure 1F). In contrast, 
2 inhibitory receptors, NKG2A and CD85j, were upregulated in the 
M group compared with control groups and the Tis group (Figure 
1, G and H). Histograms are shown in Supplemental Figure 1A (sup-
plemental material available online with this article; doi:10.1172/
JCI45816DS1). We submitted the normalized expression values of 
the 22 receptors to TMEV software analysis based on unsupervised 
hierarchical clustering. The clustering emphasized the similarities 
between the control groups and Tis tumors on one side and the 3 
invasive tumor groups on the other side (Figure 1I). The degree of 
alterations clearly progressed with tumor stage, notably in the inva-
sive tumors, with underexpression of activating receptors and over-
expression of inhibitory receptors in more advanced stages.
p-NK cell functions are altered in patients with invasive BC
We next  determined  whether  these  modifications  of  expres-
sion were associated with altered NK cell functions. We assessed 
peripheral blood NK (p-NK) cells’ functionality through their 
ability to  kill  a target cell (cytotoxic activity), to degranulate 
(CD107 positivity, associated with both  helper and cytotoxic 
functions), to produce IFN-γ and TNF-α (helper functions), and 
to mediate ADCC. We found that effective killing (Figure 2A) and 
degranulation efficiency (Figure 2B) were altered in the invasive 
groups compared with the noninvasive ones. IFN-γand TNF-α
levels were also reduced in the M group compared with the nonin-
vasive groups (Figure 2, C and D). Highly responsive NK cells can 
exert multiple functions with increased potency compared with 
monofunctional cells (33). We used the Boolean gating strategy 
to identify this multifunctional potential of NK cells. p-NK cells 
from the M group were mostly monofunctional (either CD107 
or IFN-γ or TNF-α), whereas p-NK cells from the other groups 
Table 1
Clinical features of BC patients
Groups B Tis LOC LA M
n = 19 8 55 26 32
Mean age (years) 56 51 57 52 52
Invasive margin No No Yes Yes Yes
Sample uptake Before surgery Before surgery Before surgery Before surgery Before therapy
Scarff-Bloom-Richardson (SBR) Grade I ND ND 37.0% 3.8%
II ND ND 38.8% 34.6%
III ND ND 24.0% 53.8%
pN+ No No 34.0% 81.0% 100.0%
Median pT (mm) ND 20.2 26.4 57.1 50.0
Perivascular invasion No No 16.6% 26.9% 9.3%
Lympho-plasmocytic stroma Absent ND ND 37.5% 47.7% ND
Moderate ND ND 56.2% 38.1% ND
High ND ND 6.3% 14.2% ND
IHC expression PR+ ND ND 62.9% 53.8% 46.8%
ER+ ND ND 75.9% 69.2% 40.6%
erbB-2+ ND ND 7.4% 15.3% 43.7%
HRerbB-2 ND ND 18.5% 15.3% 18.7%
Histological type Ductal ND 100% intraductal 57.4% 65.4% 82.2%
Lobular ND 18.5% 19.2% 6.2%
Tubular ND 14.8% 0.0% 0.0%
Other ND 9.2% 15.3% 12.5%
PR, progesterone receptor; ER, estrogen receptor; HR, hormone receptor, i.e., PR and /or ER+; ND, not determined.
research article
The Journal of Clinical Investigation      http://www.jci.org      Volume 121      Number 9      September 2011  3611
Figure 1
Phenotype of p-NK cell receptors in patients with dif-
ferent stages of BC at diagnosis and controls. (AH)
Receptors were significantly altered among the dif-
ferent control groups and BC patients. The individual
mean fluorescence intensities (MFI) (unimodal expres-
sion) or percentages of cells positive for given markers
(bimodal expression) were graphed for the following
receptors: (A) NKp30 (MFI); (B) NKG2D (MFI); (C)
2B4 (MFI); (D) DNAM-1 (MFI); (E) CD16 (MFI); (F)
NKp46 (MFI); (G) NKG2A (percentage of cells positive
for NKG2A); and (H) CD85j (MFI). For each scattered
plot, the horizontal bar represents the mean value. (I)
Summary of the median expression of the 22 NK cell
receptors tested in the peripheral blood of controls and
BC patients. MFI or percentage of positive cell median
values were submitted to a hierarchical clustering pro-
gram to obtain a global view of receptor expression
in the different groups. HD (n = 22), B (n = 19), Tis
(n = 7), LOC (n = 55), LA (n = 26), M (n = 32). Statisti-
cal analyses were done using nonparametric unpaired
Mann-Whitney U test and Kruskal-Wallis (KW) ANOVA.
*P < 0.05; **P0.005; ***P ≤ 0.0005.
research article
3612 The Journal of Clinical Investigation      http://www.jci.org      Volume 121      Number 9      September 2011
exerted the 3 functions simultaneously (Figure 2E). Finally, we 
measured theabilityof  p-NK  cellsto  mediate  ADCC  against 
SK-BR-3, a BC cell line sensitive to trastuzumab, an anti–erbB-2 
monoclonal antibody known to induce ADCC (3). We were not 
able to obtain enough p-NK cells to perform this test in the M 
group. In the other groups, we observed a massive enhancement 
of cytotoxicity in the presence of increasing doses of trastuzumab, 
especially in the noninvasive groups, with more than 80% of cells 
responding (Figure 2F and Supplemental Figure 2A). NK cells 
from B and Tis patients were more efficient in mediating ADCC 
at the lowest doses of trastuzumab than NK cells from LOC and 
LA patients, which is coherent with their different levels of CD16 
Figure 2
p-NK cell functions are altered in invasive BC patients. p-NK cells isolated from the different groups of patients were exposed to K562 cells in a
direct cytotoxic assay. (A) Effective killing of K562 cells. (B) Percentage of NK cells positive for CD107. (C) Percentage of NK cells positive for
IFN-γ. (D) Percentage of NK cells positive for TNF-α. (E) The multipotentiality of p-NK cells was determined from the number of functions (degranu-
lation as measured by CD107 expression, production of IFN-γ and/or TNF-α) that each p-NK cell was able to simultaneously accomplish against
K562 target cells. The E/T ratio was 1:1 in cytotoxic experiments performed against K562 cells. (F) ADCC efficiency was measured against the
SK-BR-3 BC cell line preincubated without or with increasing therapeutic doses (D1 to D7) of trastuzumab. Activation of NK cells was measured
by the expression of CD107. The E/T ratio was 2:1. The numbers of included patients per group were the following: B (n = 10), Tis (n = 7), LOC
(n = 16), LA (n = 16), M (n = 12, except for ADCC experiments where we did not obtain enough cells to perform the test). The statistical differences
between groups were established using nonparametric Mann-Whitney U test. *P < 0.05; **P0.005. Data are represented as mean ± SEM.
research article
The Journal of Clinical Investigation      http://www.jci.org      Volume 121      Number 9      September 2011  3613
expression. The lower responses observed in these patients were 
not due to a decreased proportion of the VV polymorphism, the 
CD16 allotype with the highest affinity for trastuzumab (Supple-
mental Figure 2B). To evaluate specific effects of the major pheno-
typic alterations, we performed a number of redirected cytotoxic 
assays involving NKp30 and CD16. We found that the expression
of a given receptor directly correlated with the receptor-triggered 
cytotoxicity (Supplemental Figure 2, C and D).
Thus, phenotypic alterations of p-NK cells in invasive BC corre-
late with altered functions. However, even if p-NK cells alterations 
are associated with disease progression, they might not reflect the 
situation within the tumor. We thus next studied tumor-infiltrat-
ing NK (Ti-NK) cells.
Ti-NK cells are functionally impaired compared with Mt-NK cells
Despite their major role in antitumor immunity and their presence 
in close vicinity to cancer cells (Figure 3A), the phenotype of mam-
mary-infiltrating NK cells remains unknown. To identify altera-
tions that are tumor induced rather than tissue specific, we com-
pared NK cells isolated from breast tumors (n = 24, Ti-NK) with 
paired p-NK cells (n = 11, p-NK) and/or with paired tumor-free 
NK cells isolated from symmetric normal breast samples (n= 4, 
mammary tissue–NK cells [Mt-NK]). Because of a low lymphocyte 
extraction yield, we limited our study to NK cells isolated from the 
largest LOC tumors and LA tumors. The proportion of Ti-NK cells 
was similar to that found in paired tumor-free mammary tissues 
(7.9% vs. 12.5%, respectively; Figure 3B). However, because the pro-
portion of CD45+ cells was lower in tumors than in healthy mam-
mary tissues, the absolute number of NK cells per million cells was 
severely reduced in the malignant tissues (12,320 ± 1700 vs. 43, 
265 ± 10 225 in healthy tissue; Figure 3C). Most importantly, the 
phenotype of NK cells differed in malignant and healthy tissues, at 
least with respect to their CD56Bright and CD56Dim phenotype: there 
were more CD56Bright cells infiltrating the tumor (33.9% ± 14.8%) 
than in the healthy mammary tissue (14.5% ± 6.1%). This latter 
level was more comparable to that seen in P blood (8.9 ± 6.5%) 
(Figure 3D). This increase of Ti-CD56Brightcells suggested either 
the existence of 2 different maturation stages, as CD56Bright cells 
are usually recognized as more immature NK cells with poor cyto-
toxic activity, or 2 activation stages, as the upregulation of CD56 
expression is also a characteristic of activated NK cells, indepen-
dent of their function.
To strengthen this observation, we studied other NK cell mark-
ers. The increased expression of CD56 together with NKp44 and 
CD69 (Figure 4, A and B, and Supplemental Figure 3A) suggested 
that Ti-NK cells were indeed activated NK cells. Moreover, the 
increased expression of NKG2A and CD27 and the decreased 
expression of activating receptors (NKp30, NKG2D, DNAM-1, 
CD16) and cytotoxicity-related molecules (CD57, PRF1, GZMB, 
and TRAIL) on Ti-NK cells compared with p-NK cells and Mt-
NK cells suggested that Ti-NK cells have a poor cytotoxic poten-
tial. We then submitted our NK cells’ profiles to TMEV software 
analysis. BC aggressiveness was evaluated with the Nottingham 
Prognostic Index (NPI), a prognostic index based on 3 patho-
logical parameters (tumor size, tumor grade, and axillary LN 
status) in nonmetastatic invasive BC (34). Hierarchical cluster-
ing underlined  the similarity of  the profiles between Mt-NK 
cells and p-NK cells (Figure 4C). Ti-NK cells profiles were het-
erogeneous, with the exception of those from a group of patients 
with poor prognosis (high NPI) that showed underexpression 
of nearly all NK cell receptors. NK cell markers clustered into 
3 groups: group I was composed of the cytotoxic-related effec-
tors, group II, the receptors involved in NK cell activation, and 
group III, the molecules related to NK cell maturation markers. 
Figure 3
NK cells infiltrating breast tumors
(Ti-NK cells) compared with healthy
mammary tissue (Mt-NK cells). (A)
Immunohistochemistry staining of
Ti-NK cells on a paraffin section of
a breast tumor. NK cells were iden-
tified with an anti-CD56 antibody
(brown staining). Other lympho-
cyte populations, mostly T cells,
displayed condensate dark blue
nucleus, while tumor cells are dis-
tinguishable by their larger size. (B)
FACS analysis of the percentage
of CD45+CD56+CD3 cells found
in tumor and paired healthy mam-
mary tissue. (C) Absolute number of
CD45+CD56+CD3 cells per million
cells found in tumor compared with
healthy mammary tissue. (D) Per-
centage of CD56Bright and CD56Di m
NK cells in the peripheral blood,
tumor, and healthy mammary tissue
of BC patients.
research article
3614 The Journal of Clinical Investigation      http://www.jci.org      Volume 121      Number 9      September 2011
Figure 4
Phenotype of NK cells infiltrating healthy mammary tissue (Mt-NK cells), tumors (Ti-NK cells), and comparison with peripheral blood (p-NK
cells) profile. (A) Monoparametric histograms of the most important NK cell cytotoxic receptors in paired compartments, respectively, Mt-
NK cells and Ti-NK cells. (B) Monoparametric histograms of NK cell receptors involved in NK cell maturation and/or cytotoxicity in paired
compartments, respectively, Mt-NK cells and Ti-NK cells. (C) Hierarchical cluster representation of NK cell receptors expressed on Ti-NK,
Mt-NK, and p-NK. The phenotypes of 24 Ti-NK cells, 4 paired Mt-NK, and 11 paired p-NK cells were submitted to TMEV, and data were
normalized by row; then the hierarchical clustering was applied to both NK cell markers and patient samples. Markers are represented
horizontally while each patient is graphed vertically. Patients’ NPI is shown below the clustering. The following color code was used: green,
NPI < 3.4 (good-to-excellent prognosis); yellow, 3.4 < NPI < 5.4 (moderate prognosis); blue, NPI > 5.4 (poor prognosis). (D) Result of the
contingency data for the group II metamarker. The Fisher exact test was significant (P = 0.013), and the strength of association was as
follows: relative risk: 0.1477, 95% CI: 0.02170 to 1.006; odds ratio, 0.0625; 95% CI: 0.006019 to 0.6490.
research article
The Journal of Clinical Investigation      http://www.jci.org      Volume 121      Number 9      September 2011  3615
Downregulation of group I molecules was observed in all Ti-NK 
cells compared with Mt-NK and p-NK cells, suggesting that it 
represents an alteration induced by the tumor. We searched for 
correlations among the expression of these 3 groups (gene clus-
ters) and tumor aggressiveness using a metagene approach (35). 
The metagene expression value was determined by calculating 
the mean of the normalized expression values of all genes in the 
respective gene groups (metagenes I, II, and III). The only clear 
trend that emerged was that a high expression of metagene II was 
associated with a good-to-excellent prognosis (NPI value < 3.4) 
(Figure 4D). Together, our results indicate that most Ti-NK cells 
are CD56BrightNKG2Ahi CD16loKIRlo cells  with poor cytotoxic 
potential (PRF1loGZMBloCD57lo), associated with poor outcome 
in BC. With the exception of the KIRs, Ti-NK cell alterations were 
very similar to those found in p-NK cells, but dramatically more 
pronounced. Indeed, we found positive correlations between the 
MFI of some markers expressed on Ti-NK cells with paired p-NK 
cells (NKp30 [R= 0.645, P = 0.03], CD16 [R= 0.638 P= 0.03], 
NKG2A [R = 0.774, P = 0.003], and NKG2D [R = 0.531, P = 0.07])
(Supplemental Figure 3B). The alterations detected in peripheral
blood might thus come from recirculating Ti-NK cells and/or be 
induced by soluble factors secreted by the tumor.
We next asked whether compromised Ti-NK cells are neverthe-
less able to exert a helper function, as suggested by their CD56bright
phenotype. CD107 functional assays comparing paired Ti-NK and 
p-NK cells showed that Ti-NK cells were activated more slowly 
(CD69 as early activation marker), had a decreased degranulation 
potential (CD107), and displayed strong alterations of IFN-γ and 
TNF-α production (Figure 5, A and B). Moreover, Ti-NK cells were 
unable to fulfill their ADCC function, which was not surprising 
considering the low proportion of CD16posCD56Dimcells within 
the tumor (Figure 5C).
In conclusion, we showed that NK cells are strongly altered 
within a breast tumor  compared with the equivalent tumor-
free environment. Alterations detected in peripheral blood are 
less marked but nevertheless likely to reflect what occurs in the 
Figure 5
Functionality of NK cells infiltrating tumors (Ti-NK cells) compared with paired p-NK cells. (A) Isolated NK cells from malignant tissue (red) or
peripheral blood (black) from paired sampled were used in direct cytotoxic assays on 5 paired samples from BC patients. (B) Dot plot represen-
tation of CD69, CD107, IFN-γ, and TNF-α expression in NK cells after incubation with K652 cells, according to a 1:1 E/T ratio. Isolated NK cells
were incubated overnight in medium complemented with suboptimal concentrations of IL-2 and IL-15 before incubation with K562 cells for 4
hours. (C) Dot plot representation of CD69, CD107, INF-γ, and TNF-α expression in p-NK versus Ti-NK cells after exposure to SK-BR-3 cells in
the presence of trastuzumab. The E/T ratio was of 2:1. These results were obtained on paired samples (n = 3).
research article
3616 The Journal of Clinical Investigation      http://www.jci.org      Volume 121      Number 9      September 2011
tumor. These changes are indicative of a sustained escape from 
NK cell antitumor immunity, probably induced by the malig-
nant cells. This latter assumption will be assessed below.
NK cells are modulated by the tumor microenvironment
Cancer cells express ligands for NK cell receptors. NK cell alterations 
represent a mechanism of tumor escape or tumor immuno-edit-
ing only if the affected receptors are involved in cancer cell rec-
ognition. We thus measured the expression of NK cell ligands 
on breast  tumor cells. We found that  breast tumors express 
heterogeneous levels of ligands for NK cell receptors, notably 
regarding HLA members and NKp30-ligand (L) (Figure 6,  A 
and B). Breast tumors also frequently expressed high levels of 
DNAM-1–L and NKG2D-L (Figure 6, A  and B). Most impor-
tantly, we found deregulated transcriptional expression of NK 
cell ligands in a large set of BC (n = 250) compared with healthy
mammary tissues (n = 5 pools of 5) previously profiled using 
DNA microarrays, suggesting that an editing process was indeed 
involved (Supplemental Figure 4).
Tumor cells synthesize soluble factors affecting Ti-NK cells. To deter-
mine whether  the  malignant  cells  were responsible  for  the 
observed NK  cell alterations, we exposed normal NK  cells to 
breast tumor supernatants. Tumor supernatants, but not super-
natants from normal mammary tissue, induced alterations of 
NK cell receptors concordant with  the phenotype observed on 
Ti-NK cells. The most prominent effect was observed on GZMB 
(P < 0.05), NKG2D (P < 0.05), and NKG2A (P < 0.005) (Figure 
7A). When we exposed p-NK cells to these supernatants and then 
used them in cytotoxic assays, we observed a profound inhibition 
of all NK cell functions, particularly IFN-γ secretion, compared 
with nonexposed p-NK cells (Figure 7B). These results suggest 
that soluble factors released  by tumors  (epithelial or stromal 
components) affect both NK cell phenotype and functions. We 
determined the levels of soluble factors known to be increased in 
other cancers and to alter lymphocyte functions, such as TGF-β1, 
PGE2, LGALS3, and sMICA. We also measured ADAM17, a metal-
loproteinase involved in the shedding of MICA and other mole-
cules modulating NK cell functions (Supplemental Figure 5). We 
correlated these factors with the respective expression of recep-
tors on mammary NK cells. The resulting Rcoefficients were sub-
mitted to TMEV to obtain a correlation matrix (36). The levels 
of soluble inhibitory factors positively correlated with NKG2A 
expression (Figure 7C), whereas they negatively correlated with 
molecules related to NK cell cytotoxicity (GZMB, PRF1, CD57, 
KIR, CD16). The highest negative correlations were obtained 
with TGF-β1 and PGE2. TGF-β1 also displayed the highest corre-
lation with the NPI, showing that more aggressive breast tumors 
contain more  TGF-β1–producing cells (Figure  7D). Blocking 
TGF-β1 in tumor supernatants partially restored NK cell func-
tionality, by 20% on average (Figure 7E).
Reversion of p-NK cell alterations under remission
As a final proof that NK cell alterations are tumor induced, we 
established the phenotype of p-NK cells from 7 former patients 
with invasive BC (Ex-BC) who had undergone surgery more than 
5 years ago and had not relapsed since. We matched these cases 
for age and pTNM classification with the BC patients from our 
series. Patients from the benign group (B group) were used as the 
Figure 6
Ligands of NK cell receptors are expressed by breast tumor cells.
(A) Epithelial cells isolated from malignant mammary tissue were
phenotyped for the ligands of the main altered NK cell receptors.
2 examples representative of each ligand expression are illustrat-
ed here. Horizontal bars indicate the percentage of epithelial cells
positive for the marker of interest (solid line, gray histogram) in com-
parison with the respective control isotype on the same population
(dashed line, white histogram). (B) Summary of MFI obtained for
each of the NK cells ligands in 5 independent experiments. Data are
represented as mean ± SEM.
research article
The Journal of Clinical Investigation      http://www.jci.org      Volume 121      Number 9      September 2011  3617
control for a nonmalignant situation. NKp30, CD16, NKG2D, 
and NKG2A receptors  were expressed  at similar  levels in  the B 
group and Ex-BC patients, respectively up- and downregulated 
compared with matched NK cells isolated from the BC groups, 
indicating that a normal NK cell phenotype was restored in the 
Ex-BC patients after tumor removal (Figure 8). In conclusion, our 
data show that BC cells can shape a strong inhibitory microenvi-
ronment able to inhibit NK cell antitumor immunity.
Discussion
Understanding how the immune system affects the development 
and progression of solid tumors remains one of the most challenging 
questions in immunology. It has been previously observed that mice 
with defects in IFN-γ signaling or antitumor immune cells are more 
likely to develop primary tumors (5, 37, 38). Accordingly, antitumor 
immunity is frequently deficient in patients with solid tumors (28, 29). 
We have confirmed here that p-NK cell activity, notably IFN-γ pro-
Figure 7
Impact of breast tumor stroma on p-NK cells. (A) p-NK cells were cultured for 48 hours in complete medium, dissociation supernatant from tumor-
free samples (H, n = 4), or dissociation supernatant from malignant samples (T, n = 12). The main NK cell receptors were then phenotyped. The
resulting expressions were submitted to TMEV. The significance of the observed variation was measured with a Kruskal-Wallis (KW) test. (B)
Alterations of NK cell functions were measured from 10 different tumor supernatants. CD69 expression, CD107 degranulation, IFN-γ, and TNF-α
synthesis and percentage of absolute dead K562 cells were measured from p-NK cells cultured in complete medium (white bars) and p-NK
cells exposed to tumor supernatant (gray bars). (C) The correlation matrix of the R coefficients was obtained with a Spearman’s test between
the quantitative levels of tissue-associated soluble factors and paired NK cell receptors expression. (D) Correlation between TGF-β1 levels
found in breast tumor supernatants and the respective NPI. (E) Partial restoration of NK cell functions after preincubation of tumor supernatants
(n = 10) with blocking TGF-β1 antagonist antibody. Variations were evaluated with a nonparametric Mann-Whitney U test. *P < 0.05; **P0.005;
***P ≤ 0.0005. Data are represented as mean ± SEM.
research article
3618 The Journal of Clinical Investigation      http://www.jci.org      Volume 121      Number 9      September 2011
duction, is  strongly impaired in invasive breast  tumors (39, 40), 
especially when the stage of disease is advanced. In situ, the immune 
infiltrate has long been neglected, but we now know that it is not the 
quantity but the quality of the immune infiltrate, mostly composed 
of cytotoxic cells, that can predict tumor outcome of several cancers, 
such as early-stage colorectal cancer, gastrointestinal tumor, pulmo-
nary adenocarcinoma, and, more recently, BC (41–43).
Our data show that at least 4 major mechanisms are associated 
with an escape from NK cell antitumor immunity in invasive breast 
tumors. First, we showed that breast tumor cells alter NK cell func-
tions through the modulation of their surface receptors, a mechanism 
observed in several other malignancies (28, 29, 44–47). These altera-
tions are associated with invasive characteristics and a poor progno-
sis. Among invasive BCs, the cases in which NK cell infiltrates still 
displayed high expression of NCRs, NKG2D, and DNAM-1 receptors, 
and so which could still be considered as activated cells, have a good 
prognostic index. NK cells were our primary population of interest 
considering that invasive tumors preferentially develop in an NK-defi-
cient mouse background (6, 7). However, some of the affected recep-
tors, such as DNAM-1, NKG2D, and NKG2A, are also expressed by 
specific populations of T cells. We found that these receptors are also 
altered on tumor-infiltrating T cells, but we did not detect any associa-
tion with the acquisition of invasive characteristics (data not shown). 
Data obtained on large cohorts have demonstrated a beneficial effect 
on BC patients’ survival of the presence in the tumor of cytotoxic 
T cells at diagnosis, suggesting that these cells also play a major role 
in breast tumor outcome. It would thus be interesting to look at the 
functional behavior and/or regulation of other antitumor immune 
cells. By inducing the downregulation of activating receptors and the 
upregulation of inhibitory receptors on lymphocytes, the tumor might 
become “invisible” to antitumor immunity.
Second, we  looked at the  immunogenicity  of breast tumors 
and found that both the protein and mRNA profiles of NK cells’ 
ligands are different from the profiles seen in healthy mammary 
tissues. This is also a phenomenon frequently observed in several 
tumors (28, 48, 49). The heterogeneity of the profiles suggested 
that the phenotype of tumor cells at diagnosis is the result of a 
more or less successful immuno-editing (28, 50).
Third, the alterations of Ti-NK cells, and their reversibility in 
BC patients in long-term remission, suggested that the tumor 
induces its own tolerance from NK cell antitumor immunity 
(51). We tried to identify how the tumor might influence the 
expression of activating and inhibitory ligands. Based on previ-
ously published data, we showed that the alterations of NK cell 
receptors could be induced by the breast tumor microenviron-
ment. TGF-β1 was an obvious candidate because it is a power-
ful inhibitor (52–54) elevated in advanced breast tumors (55). 
Accordingly, blocking TGF-β1 in mice suppresses the occurrence 
of metastases and restores NK cell activity (56). Furthermore, 
TGF-β1 is also one of the key factors promoting epithelial-to-
mesenchymal transition (EMT), a major mechanism active at 
the invasive front of solid tumors and responsible for disease 
progression, metastatic potential, and chemotherapeutic resis-
tance (57–60). Mesenchymal stem cells, which support cancer 
growth in situ or at the site of BC metastasis, can protect BC 
cells from immune recognition directly by producing TGF-β1 or 
indirectly by increasing the recruitment of Tregs (61, 62). This 
is in line with the correlations we observed between the in situ 
levels of TGF-β1 or Treg infiltrate and NK cells’ impaired phe-
notype and functions (refs. 31, 63, 64; Figure 7 and Supplemen-
tal Figure 6). However, because TGF-β1 only partially explained 
this effect, we did not limit our study to TGF-β1 and Tregs as 
Figure 8
Reversibility of p-NK cell phenotype in patients in remission. The phenotypes of NKp30, CD16, NKG2D, and NKG2A receptors were restored
in former BC patients compared with matched (age, group, and TNM classification) BC patients at diagnosis. BC (n = 7); B (n = 12); Ex-BC
(n = 7). Data are represented as mean ± SEM.
research article
The Journal of Clinical Investigation      http://www.jci.org      Volume 121      Number 9      September 2011  3619
potential inhibitors of NK cell functions. We found that not 
only PGE2(65, 66), sMICA (67), LGALS3 (68), and  ADAM17 
(69), but also IDO1 (data not shown and ref. 70) contributed to 
the immunosuppressive environment. All these molecules have 
previously been implicated in tumor progression and impair-
ment of NK cell phenotypes and functions. In the MMTV/Neu 
mouse model, which spontaneously develops invasive mammary 
tumors (71), we observed alterations of NK cell phenotypes and 
functions concordant with those found in humans (Supplemen-
tal Figure 7). It would be interesting to determine whether the 
administration of blocking antibodies against the various actors
of the tumor microenvironment previously mentioned, alone or 
in combination, could delay or accelerate invasive tumors and/
or metastasis occurrence in this model.
Finally, we looked at how the microenvironment influences final 
maturation of NK cells in tissues. A limited amount of informa-
tion on other tissue-infiltrating NK cell phenotypes is available 
(72, 73). NK cells usually show tissue-specific patterns of expres-
sion and disturbance of the cytotoxic CD56DimCD16+KIR+/–and 
helper/precursor CD56BrightCD16KIR balance (74, 75). Notably, 
human NK cells collected from nonreactive LNs are CD16KIR, 
whereas NK cells derived from reactive/efferent LN and blood 
express  CD56DimCD16+KIR+  (76).  The  current  belief  is  that 
immature CD56BrightNK cells acquire these molecules in the LN 
during inflammation and then circulate as KIR+CD16+NK cells. 
CD56BrightCD16KIR– and CD56DimCD16+KIR+/–NK cells might 
thus correspond to sequential steps of differentiation, like the 
CD56BrightCD16lowKIR+/– found in breast tumors (77, 78). Togeth-
er, these  observations  support  the  hypothesis that secondary 
lymphoid organs, or any microenvironment, can be the site not 
only of human NK cell final maturation but also of self-tolerance 
acquisition during immune reaction. Several recent studies show-
ing that NK cell developmental programming is not entirely fixed 
and that mature NK cells can be reeducated by their environment 
support this hypothesis (79–81). We showed that BC Ti-NK cells 
expressed activating molecules (CD56Bright, CD25, CD69, NKp44) 
like mature activated NK cells, but the levels of all other markers 
were instead characteristic of immature and nonfunctional NK 
cells (82). The most intriguing feature was the disappearance of 
the KIR, otherwise present on Mt-NK cells, confirming that Ti-
NK cells could not  exert cytotoxic functions like noneducated 
cells (83). The education process strongly depends on the engage-
ment of activating and inhibitory receptors (84), but ligands were 
extremely heterogeneous in the different breast tumors, whereas 
KIR loss was homogeneous. We suggest that a strong inhibitory 
environment can reorient or reverse the transcriptional program 
of NK cell maturation toward a nonreactive self-tolerant profile.
In conclusion, our study highlights the role of NK cells in the con-
trol of invasive breast tumors, suggesting that restoring NK cells’ 
antitumor efficiency (with adoptive transfer of preactivated or allo-
reactive NK cells) after surgery might help clear residual tumor cells 
as demonstrated in other malignancies (85–87). Most importantly, 
we showed that invasive breast tumors induce self tolerance in NK 
cells, resulting in attenuated malignant cell immunogenicity and the 
creation of a multifaceted immunosuppressive microenvironment 
that blunts NK cells’ cytotoxicity and prevents their final matura-
tion process. NK cell subversion is thus another active mechanism, 
in addition to the other breast tumor–derived mechanisms such as 
increased Treg infiltrates, dendritic cell modulation, and inhibitor 
secretion, that contributes to breast tumor  progression. Finally, 
these results are particularly interesting in the context of BC dor-
mancy and immune evasion and highlight the importance of devis-
ing future therapies able to enhance NK cell cytotoxicity efficiently 
to further prevent invasive breast tumor recurrence (88).
Methods
Patients. Benign breast tumor and BC patients treated at the Institut Paoli-
Calmettes were prospectively recruited on diagnosis between January 2007 
and December  2009. Blood and/or  tumors were respectively  sampled 
before or during the surgical diagnostic or therapeutic act, before admin-
istration of any treatment related to tumor progression. Fresh samples 
were extemporaneously treated, before the determination of the diagnosis 
by the surgeon. After analysis of morphological tumor characteristics by 
pathologists, patients were retrospectively classified into 5 groups, as fol-
lows (Table 1): patients with a benign tumor (the B group), patients with 
an in situ cancer (the Tis group), patients with invasive localized cancer 
(pT1N0 an pT2N0, referred to as the LOC group), patients with invasive 
locally advanced cancer (pT2N1-2 to pT4, referred as the LA group), and 
patients whose initial breast tumor had given rise to metastases in distant 
organ(s) (referred to as the M group). We also included a control group 
of former BC patients, followed in our institution,  who had undergone 
surgery at least 5 years ago and had not relapsed since. NPI was used to 
establish the precise prognostic value in invasive tumors, based on 3 fac-
tors: tumor grade, number of LNs involved, and size of the tumor.
NK cell phenotype by flow cytometry. 100 μl of fresh whole blood or 0.7 × 106
Ti-cells were incubated with the appropriated antibodies (Supplemen-
tal Table 1) on a rocking platform  for 30 minutes. Red blood  cells were 
lysed with OptiLyse B (Beckman  Coulter). Samples were extemporane-
ously analyzed on a BD FACS Canto (BD Biosciences). Before and after 
reading patient samples, fluorescence intensities of the FACS Canto were 
standardized over time with PMT 7-Color Setup Beads (BD Biosciences) 
to compensate for fluorescence intensity variability. The gating strategy 
consisted of the elimination of the doublets based on the FCS-A/FCS-H 
parameters, followed by the removal of dead cells. NK cell population was 
selected based on the following phenotype: CD45posCD3negCD56pos. Treg-
cells were selected on the following expression: CD45posCD3posCD4posCD
25hiCD127lo.
NK cell isolation from peripheral blood samples. NK cells were negatively iso-
lated with the NK cell StemSep System (StemCell Technology) according 
to the manufacturer’s instructions. The purity and viability of sorted cells 
were established and always greater than 94%.
NK cells were incubated in RPMI, 10% FCS, complemented with sub-
optimal concentrations of IL-2 (100 U/ml; Proleukine, Chiron) and IL-15 
(5 μg/ml; R&D Systems) overnight.
Isolation of epithelial tumor cells from mammary tissue. The minced mammary 
tumor was incubated with 1× collagenase/hyaluronidase for 16 hours at 
37°C on a  rotary shaker, according to the manufacturer’s instructions 
(StemCell Technology). Briefly, the liquefied fat layer and  supernatant 
were discarded after centrifugation. The pellet, enriched in epithelial cells, 
was resuspended in prewarmed trypsin-EDTA, and then washed. Cells were 
resuspended in prewarmed dispase (5 mg/ml) complement with DNase I 
(1 mg/ml). After additional washes, cells were counted and resuspended in 
PBS, 2% BSA, for fluorescent staining if viability exceeded 75%.
Isolation of lymphocyte infiltrates in mammary tissue. A section made within 
the core of the malignant area was selected by the pathologist and extem-
poraneously cleared of pieces of fat, then weighed; an equal volume (w/v) 
of medium (RPMI 1640) was added to the tumor. The tumor was mechan-
ically disrupted; then the supernatant was harvested and centrifuged at 
high speed to remove residual cellular elements. This supernatant, later 
referred to in the text as “supernatant of dissociation,” was aliquoted and 
research article
3620 The Journal of Clinical Investigation      http://www.jci.org      Volume 121      Number 9      September 2011
frozen at –20°C until further use. The mean volume of the tumor superna-
tant of dissociation was 0.640 ± 0.420 ml. After mechanical disruption and 
removal of the supernatant of dissociation, the tumor was digested for 1 
hour under agitation with collagenase Ia (1 mg/ml) and DNase I (50 × 103
units/ml; Sigma-Aldrich). Cell suspension was then used for flow cytom-
etry staining or NK cell isolation if viability was greater than 80%.
Cytotoxic activity analysis. NK cells were tested for cytotoxic activity against 
the leukemic HLA-Ineg K562 cell line (direct cytotoxicity), the FcγR-positive 
P815 mastocytoma murine cell line (redirected cytotoxicity). or the erbB-
2–overexpressing BC cell line SK-BR-3 (used to measure ADCC potential 
of NK cells in the presence of trastuzumab, the therapeutic mAb targeting 
erbB-2). These cytotoxic tests were all done in 4-hour assays. The measured 
parameters were degranulation (CD107a and CD107b) and cytokine pro-
duction (IFN-γand TNF-α) by NK cells or the percentage of absolute num-
ber of dead targets, referred to as the cell death index (CDI). The respective 
effector/target (E/T) ratios are indicated in the figure legends.
In redirected experiments, NK cells were preincubated (20 minutes at 
37°C) with saturating amounts of purified antibodies (IgG1, anti-CD16, 
anti-NKp46, anti-NKG2D, anti-NKG2A, anti-KIR, anti-NKp30, and anti-
DNAM-1 mAb), then used in cytotoxic assays.
In ADCC experiments, SK-BR-3 cells were preincubated (30 minutes 
at 4°C) with increasing amounts of purified trastuzumab, then used in 
cytotoxic assays.
For flow cytometric experiments measuring the degranulation and cytokine 
production of NK cells and CD107a and CD107b mAb plus GolgiStop, all 
purchased from BD Biosciences, were added in each well at the beginning of 
the culture. At the end of this incubation, NK cells were stained with CD56 
and permeabilized with Cytofix/Cytoperm (BD Biosciences). Intracellular 
antibodies were next added (IFN-γ and TNF-α) before cell analysis.
For cytotoxic assays measuring the CDI (percentage of dead targets com-
pared with targets not exposed to NK cells), experiments were done in True-
Count Beads Tubes (BD Biosciences) to precisely quantify the number of 
NK cell–mediated dead cells. Dead cells were visualized with the Live/Dead 
red reagent according to the manufacturer’s instructions (Invitrogen).
For the study of alterations induced by breast supernatants, purified NK 
cells were cultured with or without healthy or tumor supernatants (dilu-
tion 1:2) before staining 48 hours later or used in CD107 assays 40 hours 
later (followed by 4 hours of CD107 assay).
Finally, for the study of TGF-β1 involvement in breast tumor–mediated 
alterations, NK cells were cultured in the presence of breast tumor superna-
tants (dilution 1:4) preincubated with blocking anti–TGF-β1 (20 μg/ml).
Immunohistochemistry. Immunohistochemical staining was performed on 
5-μm cryostat sections fixed in 4% paraformaldehyde. After neutralization 
of the endogenous peroxidase and saturation, sections were layered with 
the anti-CD56 mAb (Dako) or isotype-matched control mAb, revealed with 
the Vectastain ABC Kit (Dako), and counterstained with Gill hematoxylin 
(Merck). Although CD56 is not strictly specific for NK cells, the percentage 
of CD56+CD3+ cells determined by flow cytometry staining was negligible. 
Therefore, we estimated that CD56 staining gave a reasonable representa-
tion of NK cell infiltrates within the tumor.
ELISA. ELISAs  were  performed  according  to  the  manufacturer’s 
instructions. sMICA, TGF-β1, LGALS3, and ADAM17 were purchased 
from R&D Systems, while the PGE2 ELISA kit was purchased from Cay-
man Chemical (Interchim).
Microarray experiments.  U133+2 pan-genomic Affymetrix data from 
malignant and healthy mammary tissues were downloaded from the public 
GEO data sets (GSE21653).
MMTV-Neu mice. FVB/N-Tg(MMTVneu)202 Mul/J mice (MMTV-neu 
mice) were obtained  from the Jackson Laboratory. All  mice were main-
tained on an FVB background. Mice were palpated weekly from 5 months 
of age and, once mammary gland nodules were detected, tumor diameter 
was measured twice weekly with calipers. MMTV-Neu mice (n= 6) and FVB 
control mice (n= 6) were compared, and a Kaplan-Meier analysis was done 
to determine tumor-free survival. Monthly blood samples were collected 
under isoflurane (5%) inhalation anesthesia and analyzed by f low cytom-
etry. At the end of the protocol, mice were sacrificed and breast tumors and 
organs were collected.
Statistics. All statistical analyses were done with StatView software. Dif-
ferences between breast tumor patients’ groups were evaluated using 
2-tailed nonparametric unpaired Mann-Whitney Utest. The comparisons 
between NK cell markers in paired tumor and peripheral blood samples 
were done with nonparametric Wilcoxon tests. Finally, all correlations 
between MFIs of NK cell markers expressed  on Ti-NK and p-NK were 
evaluated with nonparametric Spearman correlation tests. The equality 
of the population  medians among the different groups  was triggered 
with the nonparametric ANOVA (Kruskal-Wallis) for each marker. In all 
figures with histograms, data are  represented using the mean ± SEM. 
P < 0.05 was considered as significant.
Human study approval. This study was approved by the Institutional Ethic 
Committee Review Board (Comité d’orientation Stratégique [COS], Mar-
seille, France) from the Institut Paoli-Calmettes. During the inclusion visit, 
each patient gave writteninformed consent for research use.
Mouse study approval. All animal procedures were performed in accordance 
with protocols approved by the local Committee for Animal Experimenta-
tion (CAE) of Marseille, France.
Acknowledgments
We thank the Departments of Medical Oncology, Surgical Oncol-
ogy, Pathology and the crew of the “Hôpital de jour” at the Insti-
tut Paoli-Calmettes as well as the patients, and controls for their 
contribution to this work. We also want to thank J. Ewbank for 
careful editing of the manuscript. The “Institut National du Can-
cer” and ANR-RIB funded this study. E. Mamessier was funded 
by the Association pour la Recherche contre le Cancer for 3 years. 
A. Moretta was funded by AIRC: IG project n. 10643 and Special 
Project 5x1000 n. 9962.
Received for  publication November 17,  2010, and accepted  in 
revised form June 29, 2011.
Address correspondence  to: Emilie Mamessier  or Daniel Olive, 
INSERM UMR  891, 27 Bd  Lei Roure, 13009  Marseille, France. 
Phone: 33.491.758.415; Fax: 33.491.260.364; E-mail: mamessier@
ciml.univ-mrs.fr (E. Mamessier), daniel.olive@inserm.fr (D. Olive).
Emilie Mamessier’s present address is: Laboratory Genomic Insta-
bility and Human Hemopathies, CIML, Parc Scientifique de Lumi-
ny, 13288 Marseille, France.
1. Tirona  MT, Sehgal  R,  Ballester  O.  Prevention 
of breast  cancer (part  I): epidemiology,  risk fac-
tors,  and  risk  assessment  tools.  Cancer Invest. 
2010;28(7):743–750.
  2. Joyce JA, Pollard JW. Microenvironmental regulation 
of metastasis. Nat Rev Cancer. 2009;9(4):239–252.
  3. Kute  TE, et  al. Breast  tumor cells isolated  from 
in vitro resistance to trastuzumab  remain sensi-
tive to  trastuzumab  anti-tumor  effects  in  vivo 
and to ADCC killing. Cancer Immunol Immunother. 
2009;58(11):1887–1896.
  4. Smyth MJ, Dunn GP, Schreiber RD. Cancer immu-
nosurveillance and  immunoediting: the  roles of 
immunity in suppressing tumor development and 
shaping tumor immunogenicity.  Adv Immunol. 
2006;90:1–50.
5. Sha nkara n  V,  et  a l. IFN-gamma  and  lympho-
cytes prevent  primary  tumor  development and 
sha pe  tum or  imm unoge nicit y.  Nature.  200 1; 
410(6832):1107–1111.
  6. Dewan MZ, et  al.  Natural-killer-cells  in breast 
research article
The Journal of Clinical Investigation      http://www.jci.org      Volume 121      Number 9      September 2011  3621
cancer cell growth and metastasis in  SCID mice. 
Biomed Pharmacother. 2005;59 suppl 2:S375–S379.
  7. Dewan MZ,  et al.  Role of  Natural-killer-cells in 
hormone-independent rapid tumor formation and 
spontaneous metastasis  of breast cancer cells in 
vivo. Breast Cancer Res Treat. 2007;104(3):267–275.
  8. Olkhanud PB, et  al.  Breast  cancer lung metas-
tasis requires  expression of  chemokine receptor 
CCR4 and  regulatory  T  cells.  Cancer Res.  2009; 
69(14):5996–6004.
  9. Treilleux I,  et al. Dendritic  cell infiltration and 
prognosis of early stage breast cancer. Clin Cancer
Res. 2004;10(22):7466–7474.
  10. Gob ert  M,  et   al. Regulatory  T  cells  recruited 
through CCL22/CCR4 are selectively activated in 
lymphoid infiltrates surrounding primary breast 
tumors and lead  to an adverse clinical outcome. 
Cancer Res. 2009;69(5):2000–2009.
  11. Caras I, et al. Evidence for immune defects in breast 
and lung cancer patients. Cancer Immunol Immuno-
ther. 2004;53(12):1146–1152.
  12. Alexe G, et al. High expression of lymphocyte-asso-
ciated genes in node-negative HER2+ breast can-
cers correlates with lower recurrence rates. Cancer
Res. 2007;67(22):10669–10676.
  13. Bertucci F,  et al. Gene expression profiling shows 
medullary breast  cancer  is  a subgroup  of  basal 
breast cancers. Cancer Res. 2006;66(9):4636–4644.
14. Fin ak  G,  e t  al. Stromal  gene  expression  pre-
dicts clinical outcome  in breast cancer. Nat Med. 
2008;14(5):518–527.
  15. Sabatier R, et al. A gene expression signature identi-
fies two prognostic subgroups of basal breast can-
cer. Breast Cancer Res Treat. 2011;126(2):407–420.
16. Trinchieri G. Biology of Natural-killer-cells. Adv
Immunol. 1989;47:187–376.
  17. Vivier E, Tomasello E, Baratin M, Walzer T, Ugolini 
S. Functions of Natural-killer-cells. Nat Immunol. 
2008;9(5):503–510.
  18. Morett a  L,  Bottin o  C,  Pende  D,  Vi tale  M,  Mi n-
gari MC, Moretta  A.  Different  checkpoints  in 
human NK-cell activation. Trends Immunol. 2004; 
25(12):670–676.
19. Lanier LL. NK cell  receptors. Annu Rev Immunol. 
1998;16:359–393.
  20. Moretta A, et al. Activating receptors and corecep-
tors involved in human Natural-killer cell-mediat-
ed cytolysis. Annu Rev Immunol. 2001;19:197–223.
  21. Bauer S, et al. Activation of NK-cells and T cells by 
NKG2D, a receptor for stress-inducible MICA. Sci-
ence. 1999;285(5428):727–729.
  22. Moret ta  L,  B ottin o  C,  Pe nde  D,   Castr icon i  R, 
Mingari MC,  Moretta  A.  Surface  NK  receptors 
and their ligands on  tumor cells. Semin Immunol. 
2006;18(3):151–158.
  23. Moretta  A , et  al . Receptors for  HLA-class I-mol-
ecules in human  Natural-killer-cells.  Annu Rev
Immunol. 1996;14:619–648.
  24. Moretta A, et al. Existence of both inhibitory (p58) 
and activatory (p50)  receptors  for HLA-C mol-
ecules in  human  natural-killer cells. J Exp Med. 
1995;182(3):875–884.
 25. Long  EO.  Regulation  of  immune  response 
through inhibitory receptors. Annu Rev Immunol. 
1999;17:875–904.
  26. Lopez-Botet M, et al. Structure and function of the 
CD94 C-type lectin receptor complex involved in 
the recognition of HLA class I molecules. Immunol
Rev. 1997;155:165–174.
  27. Madjd Z, Spendlove I, Pinder SE, Ellis IO, Durrant 
LG. Total loss of  MHC class-I  is an independent 
indicator of good prognosis in breast cancer. Int J
Cancer. 2005;117(2):248–255.
  28. McGilvray RW, et al. NKG2D ligand expression in 
human colorectal cancer reveals associations with 
prognosis and evidence  for immunoediting. Clin
Cancer Res. 2009;15(22):6993–7002.
  29. Le Mau x Chan sac B,  et  al. NK-cells infiltrating  a 
MHC class-I-def icient  lung  adenocarcinoma dis-
play impaired cytotoxic activity toward autologous 
tumor cells associated with altered NK-cell-trigger-
ing receptors. J Immunol. 2005;175(9):5790–5798.
  30. Della Chiesa M, et al. The tryptophan catabolite L-
kynurenine inhibits the surface expression of NKp46- 
and NKG2D-activating receptors and regulates NK-
cell function.Blood. 2006;108(13):4118–4125.
  31. Castric oni  R,  et  al. Transforming growth factor 
beta 1 inhibits expression of NKp30 and NKG2D 
receptors: Consequences for  the  NK-mediated 
killing of dendritic cells. Proc Natl Acad Sci U S A. 
2003;100(7):4120–4125.
32. Yarbro JW, Page  DL,  Fielding LP,  Partridge  EE, 
Murphy GP. American Joint Committee on Cancer 
prognostic factors consensus conference.  Cancer. 
1999;86(11):2436–2446.
  33. Bryceson  YT, et  al. Functional analysis of human 
NK-c ells  b y  fl ow  cyto metr y.  Methods Mol Biol. 
2010;612:335–352.
 34. Todd  JH,  et  al.  Confirmation  of  a  prognos-
tic index  in  primary  breast  cancer. Br J Cancer. 
1987;56(4):489–492.
  35. Rosenwal d A , et  al . The use of molecular  profil-
ing to predict survival after chemotherapy for dif-
fuse large-B-cell  lymphoma. N Engl J Med.  2002; 
346(25):1937–1947.
  36. Galon J, et al. Type, density, and location of immune 
cells within human colorectal tumors predict clini-
cal outcome. Science. 2006;313(5795):1960–1964.
  37. Dighe AS,  Richards  E,  Old  LJ,  Schreiber  RD. 
Enhanced in vivo growth and resistance to rejection 
of tumor cells expressing  dominant negative IFN 
gamma receptors. Immunity. 1994;1(6):447–456.
  38. Kaplan DH, et al. Demonstration of an interferon 
gamma-dependent tumor sur veillance system in 
immunocompetent mice. Proc Natl Acad Sci U S A. 
1998;95(13):7556–7561.
  39. Garner WL, Minton JP, James AG, Hoffmann CC. 
Human breast cancer and impaired NK-cell func-
tion. J Surg Oncol. 1983;24(1):64–66
  40. Dewan MZ, et al. Natural-killer activity of periph-
eral-blood  mononuclear  cells  in  breast  cancer 
patients. Biomed Pharmacother. 2009;63(9):703–706.
  41. Pagès F, et al. In situ cytotoxic and memory T cells 
predict outcome in patients with early-stage colorec-
tal cancer. J Clin Oncol. 2009;27(35):5944–5951.
  42. Takanami I, Takeuchi  K, Giga  M. The pro gnostic 
value of Natural-killer cell infiltration in resected 
pulmonary adenocarcinoma.  J Thorac Cardiovasc
Surg. 2001;121(6):1058–1063.
  43. Mahmoud SM, et al. Tumor-infiltrating CD8+ lym-
phocytes predict clinical outcome in breast cancer. 
J Clin Oncol. 2011;29(15):1949–1955.
  44. Costello R T, et al. Defective expression and func-
tion   of  Nat ural- kille r  cel l-tri ggeri ng  rec eptor s 
in patients  with acute myeloid  leukemia. Blood. 
2002;99(10):3661–3667.
  45. Lakshmikanth T, et al. NCRs and DNAM-1 medi-
ate NK-cell recognition  and lysis  of human and 
mouse melanoma cell lines  in-vitro and  in-vivo. 
J Clin Invest. 2009;119(5):1251–1263.
  46. Epling-Burnette PK, et al. Reduced Natural-killer 
function associated with high-risk myelodysplas-
tic syndrome and reduced expression of activating 
NK-receptors. Blood. 2007;109(11):4816–4824.
  47. Carlsten M, et al. Reduced DNAM-1 expression on 
bone marrow NK-cells  associated with  impaired 
killing of  CD34+ blasts  in myelodysplastic syn-
drome. Leukemia. 2010;24(9):1607–1616.
  48. Bedel R , et  al. Novel  role  for STAT3 in  transcrip-
tio nal  re gulat ion  of   NK-im mune  c ell  ta rget-
ing  r ecept or  MIC A  on  c ancer   cells .  Cancer Res. 
2011;71(5):1615–1626.
  49. Kloess S, et al. IL-2-activated  haploidentical NK-
cells restore NKG2D-mediated NK-cell cytotoxicity 
in neuroblastoma patients by scavenging of plasma 
MICA. Eur J Immunol. 2010;40(11):3255–3267.
 50. Swann JB,  Smyth  MJ.  Immune  surveillance  of 
tumors. J Clin Invest. 2007;117(5):1137–1146.
  51. Fauriat  C,  et  al. Deficient expression  of NCR  in 
NK-cells from acute myeloid leukemia: Evolution 
during leukemia treatment and  impact of leuke-
mia cells in  NCRdull phenotype induction. Blood. 
2007;109(1):323–330.
 52. Flavell RA,  Sanjabi  S,  Wrzesinski  SH,  Licona-
Limón P. The polarization of immune cells in the 
tumor environment by TGF-beta. Nat Rev Immunol. 
2010;10(8):554–567.
  53. Yang L, Pang Y, Moses HL. TGF-beta and immune 
cells: an important regulatory axis in the  tumor 
microenvironment and progression. Trends Immu-
nol. 2010;31(6):220–227.
  54. Toomey  D, et  al. TGF-beta1 is elevated  in breast 
cancer tissue  and regulates  nitric oxide  produc-
tion from  a  number  of cellular sources  during 
hypoxia re-oxygenation injury. Br J Biomed Sci. 2001; 
58(3):177–183.
  55. Arteaga CL, Hurd SD, Winnier AR, Johnson MD, 
Fendly BM, Forbes JT. Anti-transforming growth 
factor-beta  antibodie s inhibit breast ca ncer  cell 
tumorigenicity and increase mouse  spleen Natu-
ral-killer-cell activity. Implications for a possible 
role of tumor cell/host  TGF-beta  interactions 
in human breast cancer  progression. J Clin Invest. 
1993;92(6):2569–2576.
  56. Thiery JP, Sl eeman JP. Complex  networks orch es-
trate epithelial-mesenchymal transitions. Nat R ev
Mol Cell Biol. 2006;7(2):131–142.
  57. Christfori G. New signals from the invasive front. 
Nature. 2006;441(7092):444–450.
  58. Brabletz T, Jung A, Spaderna S,  Hlubek F, Kirch-
ner T. Migrating cancer stem cells- an integrated 
concept of malignant tumor progression. Nat R ev
Cancer. 2005;5(9):744–749.
  59. Yang J, Weinberg RA. Epithelial-mesenchymal tran-
sition: at the crossroads of development and tumor 
metastasis. Dev Cell. 2008;14(6):818–829.
  60. Santisteban M, et al. Immune-induced epithelial to 
mesenchymal transition in vivo generates breast can-
cer stem cells. Cancer Res. 2009;69(7):2887–2895.
  61. Patel SA, Meyer  JR, Greco SJ, Co rcoran KE, Bryan 
M, Rameshwar  P. Mesenchymal  stem  cells pro-
tect breast cancer cells through regulatory T cells: 
role of mesenchymal stem cell-derived TGF-beta. 
J Immunol. 2010;184(10):5885–5894.
  62. Lin SY, et  al. The isolation of  novel mesenchymal 
stromal cell  chemotactic  factors  from the con-
ditioned  medium  of  tumor  cells.  Exp Cell Res. 
2008;314(17):3107–3117.
  63. Lee JC, Lee  KM, Kim DW, Heo DS. Elevated TGF-
beta1 secretion and down-modulation of NKG2D 
underlies  impaired  NK  cytotoxicity in  cancer 
patients. J Immunol. 2004;172(12):7335–7340.
64. Smyth MJ, Teng MW, Swann J, Kyparissoudis  K, 
Godfrey DI, Hayakawa Y. CD4+CD25+ T regulato-
ry cells suppress NK-cell-mediated immunotherapy 
of cancer. J Immunol. 2006;176(3):1582–1587.
  65. Martinet L, Jean C, Dietrich G, Fournié JJ, Poupot 
R. PGE2  inhibits  Natural-killer  and  gammad-
elta T-cell cytotoxicity triggered by NKR and TCR 
through a cAMP-mediated PKA type I-dependent 
signaling. Biochem Pharmacol. 2010;80(6):838–845.
  66. Kundu N, Ma X, Holt D, Goloubeva  O, Ostrand-
Rosenberg  S,  Fulton  AM. Antagonism  of  the 
prostaglandin E receptor EP4 inhibits metastasis 
and enhances NK function. Breast Cancer Res Treat. 
2009;117(2):235–242.
  67. Madjd  Z,  et a l. Upregulation of  MICA on high-
grade invasive operable breast carcinoma. Cancer
Immun. 2007;7:17.
  68. Demotte N,  et al. A  galectin-3 ligand corrects the 
impaired function of human CD4 and CD8 tumor-
infiltrating lymphocytes and favors tumor rejec-
tion in mice. Cancer Res. 2010;70(19):7476–7488.
  69. McGowan PM, Ryan  BM, Hill AD,  McDermott E, 
research article
3622 The Journal of Clinical Investigation      http://www.jci.org      Volume 121      Number 9      September 2011
O’Higgins N, Duffy MJ. ADAM-17 expression  in 
breast cancer correlates with variables of tumor pro-
gression. Clin Cancer Res. 2007;13(8):2335–2343.
  70. Godin-Ethier J, et al. Human activated T-lymphocytes 
modulate IDO expression in tumors through Th1/
Th2 balance. J Immunol. 2009;183(12):7752–7760.
  71. Guy CT, Webster MA, Schaller M, Parsons TJ, Car-
diff RD, Muller WJ. Expression of the neu-protoon-
cogene in the mammary epithelium of transgenic 
mice induces metastatic disease. Proc Natl Acad Sci
U S A. 1992;89(22):10578–10582.
  72. Luci  C,  et  al. Influence of  the  transcription fac-
tor RORgammat on the development of NKp46+ 
cell populations  in  gut  and skin.  Nat Immunol. 
2009;10(1):75–82.
  73. Castriconi R, et al. Natural-killer cell-mediated kill-
ing of freshly isolated neuroblastoma cells: critical 
role of DNAX accessory molecule-1-poliovirus recep-
tor interaction. Cancer Res. 2004;64(24):9180–9184.
  74. Cooper MA, Fehniger TA , Caligiuri MA. The  biol-
ogy of  human Natural-killer-cell subsets.  Trends
Immunol. 2001;22(11):633–640.
  75. Ferlazzo G, et al. Distinct roles of IL-12 and IL-15 
in human Natural-killer cell activation by dendrit-
ic-cells from secondary lymphoid organs. Proc Natl
Acad Sci U S A. 2004;101(47):16606–16611.
  76. Romagnani C, et al. CD56brightCD16– killer Ig-like 
receptor- NK-cells display  longer telomeres and 
acquire features of CD56dim NK-cells upon activa-
tion. J Immunol. 2007;178(8):4947–4955.
  77. Freud AG, Caligiuri MA. Human Natural-killer cell 
development. Immunol Rev. 2006;214:56–72.
  78. Freud   AG,  et  a l. Evidence  for discrete stages of 
human Natural-killer cell differentiation  in vivo. 
J Exp Med. 2006;203(4):1033–1043.
  79. Sun JC, Beilke JN, Bezman NA, Lanier LL. Homeo-
static proliferation generates long-lived Natural-
killer-cells that  respond against viral  infection. 
J Exp Med. 2011;208(2):357–368.
  80. Joncker NT, Shifrin N, Delebecque F, Raulet DH. 
Mature Natural-killer-cells reset their responsive-
ness when  exposed to  an altered  MHC environ-
ment. J Exp Med. 2010;207(10):2065–2072.
  81. Elliott JM, Wahle JA, Yokoyama WM. MHC class-I-
deficient Natural-killer-cells acquire a licensed phe-
notype after transfer into an MHC class I-sufficient 
environment. J Exp Med. 2010;207(10):2073–2079.
82. Carre ga  P,  et al. Natural-killer-cells  infiltrating 
human non-small-cell lung cancer are enriched 
in CD56BrightCD16 cells and display an impaired 
capability  to  kill  tumor  cells.  Cancer.  2008; 
112(4):863–875.
83. Jonsson AH,  Yokoyama WM. Natural-killer cell 
tolerance licensing  and  other mechanisms.  Adv
Immunol. 2009;101:27–79.
  84. Sun JC. Re-educating Natural-killer-cells. J Exp Med. 
2010;207(10):2049–2052.
 85. Berg M,  et  al.  Clinical-grade  ex  vivo-expanded 
human Natural-killer-cells  up-regulate  activat-
ing receptors and death receptor ligands and have 
enhanced cytolytic  activity  against tumor  cells. 
Cytotherapy. 2009;11(3):341–355.
  86. Romagné  F,  et   al. Preclinical  characterization  of 
1-7F9, a novel human anti-KIR receptor therapeutic 
antibody that augments Natural-killer-mediated kill-
ing of tumor cells. Blood. 2009;114(13):2667–2677.
87. Ru gg er i  L ,  et  al . Donor  Natural-killer  cell 
allorecognition of missing-self  in haploidentical 
hematopoietic transplantation for acute  myeloid 
leukemia: challenging its  predictive value. Blood. 
2007;110(1):433–440.
 88. Kundu  N,  Walser  TC,  Ma  X,  Fulton  AM.
Cyclooxygenase inhibitors modulate NK activities 
that control  metastatic disease. Cancer Immunol
Immunother. 2005;54(10):981–987.
... [13,14] Supporting this, NK cell deficiency by the antibody-mediated depletion or genetic targeting of NKp46+ cells leads to marked increases in metastatic colonization but not primary tumor growth in experimental mouse models. [15,16] In addition, impaired NK cell cytotoxic activity has been observed in circulation and TME during metastatic progression [2,[17][18][19] and correlates with metastatic burden in diverse cancer types. [20,21] Mechanistic studies have associated the metastatic subversion of NK cell surveillance with altered NK cell phenotypes, characterized by the downregulation of activating receptors or upregulation of inhibitory receptors, along with impaired cytotoxic potential. ...
... [20,21] Mechanistic studies have associated the metastatic subversion of NK cell surveillance with altered NK cell phenotypes, characterized by the downregulation of activating receptors or upregulation of inhibitory receptors, along with impaired cytotoxic potential. [2,17,19] Moreover, recent studies have revealed functional defects of patient NK cells without cognate receptor alterations in the periphery, [18,22,23] suggesting the contribution of distinct mechanisms to NK cell dysfunction beyond phenotypic alteration. However, few studies have been conducted on the cell-intrinsic molecular mechanism governing NK cell dysfunction that occurs during spontaneous metastasis formation and its causal contribution to metastatic outcomes. ...
Article
Full-text available
Distant metastasis, the leading cause of cancer death, is efficiently kept in check by immune surveillance. Studies have uncovered peripheral natural killer (NK) cells as key antimetastatic effectors and their dysregulation during metastasis. However, the molecular mechanism governing NK cell dysfunction links to metastasis remains elusive. Herein, MAP4K1 encoding HPK1 is aberrantly overexpressed in dysfunctional NK cells in the periphery and the metastatic site. Conditional HPK1 overexpression in NK cells suffices to exacerbate melanoma lung metastasis but not primary tumor growth. Conversely, MAP4K1‐deficient mice are resistant to metastasis and further protected by combined immune‐checkpoint inhibitors. Mechanistically, HPK1 restrains NK cell cytotoxicity and expansion via activating receptors. Likewise, HPK1 limits human NK cell activation and associates with melanoma NK cell dysfunction couples to TGF‐β1 and patient response to immune checkpoint therapy. Thus, HPK1 is an intracellular checkpoint controlling NK‐target cell responses, which is dysregulated and hijacked by tumors during metastatic progression.
... For example, Glasner et al. demonstrated that the activation of the NK natural cytotoxic receptor 1(Ncr1) (mouse) and NKp46 (human) modulated fibronectin 1 expression on tumor cells by triggering IFNγ production, consequently preventing tumor metastasis (19). Several studies have observed decreased expression of activating receptors in peripheral NK cells from patients with breast cancer, NSCLC, neuroblastoma or gastrointestinal stromal tumors, compared with healthy samples, which revealed the impaired cytotoxicity and specificity of NK cells in the tumor microenvironment (20)(21)(22)(23)(24). Therefore, we explored the NK cell perturbations in the tumor microenvironment across 26 cancer types in different sex and age groups (Fig. 4). ...
Preprint
Full-text available
Understanding how cancer immunoediting sculpts tumor microenvironments is essential to disentangling tumor immune evasion mechanisms and developing immunotherapies. Here, we construct a comprehensive immunoediting map of human cancers via single-cell deconvolution of 11057 tumor-derived samples across 33 cancer types from TCGA and comparison with 17382 healthy samples across 30 tissues from GTEx. The map covers >1000 different cell states across all the major immune cell types. Mast cells, megakaryocytes, macrophages, neutrophils, plasma cells, and T cells are up-regulated across a wide range of tumor types while natural killer cells and platelets are down-regulated in most tumor types, suggesting common cancer immunoediting events. While tumor heterogeneity is higher than the normal corresponding tissues, significant immune homogeneity exists among different tumor types compared with the distinct immune composition among normal tissues and organs. Our study provides a new holistic perspective to understanding cancer immunoediting. Our findings may provide important hints for developing novel cancer immunotherapies, and the high-resolution immunoediting map may serve as a rich resource for further pan-cancer investigation.
... The action of these activating and adhesion receptors likely counteracts the expression of NKG2A, a receptor that associates with CD94 to form heterodimers with inhibitory function 39,40 . NKG2A is currently regarded as a checkpoint receptor that decreases the ability of NK cells to recognize tumour cells, leading to lower cytotoxic function and IFNγ production [41][42][43][44] . Nevertheless, our data confirm previous results showing that NKG2D ligation can transduce a dominant stimulatory signal to NK cells, overcoming an MHC class I-mediated inhibitory signal and triggering NK cytotoxicity 45,46 . ...
Article
Full-text available
The short-lived nature and heterogeneity of Natural Killer (NK) cells limit the development of NK cell-based therapies, despite their proven safety and efficacy against cancer. Here, we describe the biological basis, detailed phenotype and function of long-lived anti-tumour human NK cells (CD56highCD16⁺), obtained without cell sorting or feeder cells, after priming of peripheral blood cells with Bacillus Calmette-Guérin (BCG). Further, we demonstrate that survival doses of a cytokine combination, excluding IL18, administered just weekly to BCG-primed NK cells avoids innate lymphocyte exhaustion and leads to specific long-term proliferation of innate cells that exert potent cytotoxic function against a broad range of solid tumours, mainly through NKG2D. Strikingly, a NKG2C⁺CD57⁻FcεRIγ⁺ NK cell population expands after BCG and cytokine stimulation, independently of HCMV serology. This strategy was exploited to rescue anti-tumour NK cells even from the suppressor environment of cancer patients’ bone marrow, demonstrating that BCG confers durable anti-tumour features to NK cells.
... As a result, the surface expression of CD4 antigen on lymphocytes decreases and the reactivity of CD8 + T cells increases. However, during this process, the body is in an immunosuppressed state, and the increased number of CD8 + T cells is dominated by Ts cells, further exacerbating the suppressive effects on CD4 + T cells 53,54 . Therefore, an increase in CD8 + T cells proportion and a decrease in CD4 + /CD8 + ratio were observed after NAT among patients with high tumor burden in the non-pCR group. ...
Article
Full-text available
Lymphocyte subsets are the most intuitive expression of the body’s immune ability, and the lymphocyte-to-monocyte ratio (LMR) also clearly reflect the degree of chronic inflammation activity. The purpose of this study is to investigate their predictive value of lymphocyte subsets and LMR to neoadjuvant therapy (NAT) efficacy in breast cancer patients. In this study, lymphocyte subsets and LMR were compared between breast cancer patients (n = 70) and benign breast tumor female populations (n = 48). Breast cancer patients were treated with NAT, and the chemotherapy response of the breast was evaluated using established criteria. The differences in lymphocyte subsets and LMR were also compared between pathological complete response (pCR) and non-pCR patients before and after NAT. Finally, data were analyzed using SPSS. The analytical results demonstrated that breast cancer patients showed significantly lower levels of CD3 + T cells, CD4 + T cells, CD4 + /CD8 + ratio, NK cells, and LMR compared to benign breast tumor women (P < 0.05). Among breast cancer patients, those who achieved pCR had higher levels of CD4 + T cells, NK cells, and LMR before NAT (P < 0.05). NAT increased CD4 + /CD8 + ratio and decreased CD8 + T cells in pCR patients (P < 0.05). Additionally, both pCR and non-pCR patients exhibited an increase in CD3 + T cells and CD4 + T cells after treatment, but the increase was significantly higher in pCR patients (P < 0.05). Conversely, both pCR and non-pCR patients experienced a decrease in LMR after treatment. However, this decrease was significantly lower in pCR patients (P < 0.05). These indicators demonstrated their predictive value for therapeutic efficacy. In conclusion, breast cancer patients experience tumor-related immunosuppression and high chronic inflammation response. But this phenomenon can be reversed to varying degrees by NAT. It has been found that lymphocyte subsets and LMR have good predictive value for pCR. Therefore, these markers can be utilized to identify individuals who are insensitive to NAT early on, enabling the adjustment of treatment plans and achieving precise breast cancer treatment.
... 2021). BC involves a higher percentage of NK2-cells, having greater activation biomarker expression, such as CD69 and NKP44, which could be due to interaction with tumour cells (Mamessier et al., 2011a;Ostapchuk et al., 2015;Mamessier et al., 2013). Ti-NK cells' anti-tumour activity and cytotoxicity decrease in the BC-related TME, having an inhibitory and exhausted phenotype associated with attenuated cytotoxic function (Mamessier et al., 2011b). ...
... IL-2 deprivation likely occurs during NK cell adoptive therapy without IL-2 support in vivo or in the TME, partly due to regulatory T (Treg) cells (55,56). This discordant regulation between NKp30 and NKp46 by IL-2 deprivation is compatible with a recent study on the sustained expression of NKp46 but the downregulation of other activating receptors, such as NKp30, NKG2D, and NKp44, in tumor-infiltrating NK cells in diverse leukemia and tumors (49,57,58). In this regard, we do not exclude alternative mechanisms to regulate the expression of N K p 3 0 c o m p a r e d t o t h a t o f N K p 4 6 , w h i c h m e r i t s further investigation. ...
Article
Full-text available
Natural killer (NK) cells are key effectors in cancer immunosurveillance, eliminating a broad spectrum of cancer cells without major histocompatibility complex (MHC) specificity and graft-versus-host diseases (GvHD) risk. The use of allogeneic NK cell therapies from healthy donors has demonstrated favorable clinical efficacies in treating diverse cancers, particularly hematologic malignancies, but it requires cytokines such as IL-2 to primarily support NK cell persistence and expansion. However, the role of IL-2 in the regulation of activating receptors and the function of NK cells expanded for clinical trials is poorly understood and needs clarification for the full engagement of NK cells in cancer immunotherapy. Here, we demonstrated that IL-2 deprivation significantly impaired the cytotoxicity of primary expanded NK cells by preferentially downregulating NKp30 but not NKp46 despite their common adaptor requirement for expression and function. Using NK92 and IL-2-producing NK92MI cells, we observed that NKp30-mediated cytotoxicity against myeloid leukemia cells such as K562 and THP-1 cells expressing B7-H6, a ligand for NKp30, was severely impaired by IL-2 deprivation. Furthermore, IL-2 deficiency-mediated NK cell dysfunction was overcome by the ectopic overexpression of an immunostimulatory NKp30 isoform such as NKp30a or NKp30b. In particular, NKp30a overexpression in NK92 cells improved the clearance of THP-1 cells in vivo without IL-2 supplementation. Collectively, our results highlight the distinct role of IL-2 in the regulation of NKp30 compared to that of NKp46 and suggest NKp30 upregulation, as shown here by ectopic overexpression, as a viable modality to harness NK cells in cancer immunotherapy, possibly in combination with IL-2 immunocytokines.
Article
Full-text available
The need for reliable biomarkers to predict clinical benefit from anti-PD1 treatment in metastatic melanoma (MM) patients remains unmet. Several parameters have been considered in the tumor environment or the blood, but none has yet achieved sufficient accuracy for routine clinical practice. Whole blood samples from MM patients receiving second-line anti-PD1 treatment (NCT02626065), collected longitudinally, were analyzed by flow cytometry to assess the immune cell subsets absolute numbers, the expression of immune checkpoints or ligands on T cells and the functionality of innate immune cells and T cells. Clinical response was assessed according to Progression-Free Survival (PFS) status at one-year following initiation of anti-PD1 (responders: PFS > 1 year; non-responders: PFS ≤ 1 year). At baseline, several phenotypic and functional alterations in blood immune cells were observed in MM patients compared to healthy donors, but only the proportion of polyfunctional memory CD4+ T cells was associated with response to anti-PD1. Under treatment, a decreased frequency of HVEM on CD4+ and CD8+ T cells after 3 months of treatment identified responding patients, whereas its receptor BTLA was not modulated. Both reduced proportion of CD69-expressing CD4+ and CD8+ T cells and increased number of polyfunctional blood memory T cells after 3 months of treatment were associated with response to anti-PD1. Of upmost importance, the combination of changes of all these markers accurately discriminated between responding and non-responding patients. These results suggest that drugs targeting HVEM/BTLA pathway may be of interest to improve anti-PD1 efficacy.
Article
Full-text available
Natural killer (NK) cells, as innate lymphocytes, possess cytotoxic capabilities and engage target cells through a repertoire of activating and inhibitory receptors. Particularly, natural killer group 2, member D (NKG2D) receptor on NK cells recognizes stress‐induced ligands—the MHC class I chain‐related molecules A and B (MICA/B) presented on tumor cells and is key to trigger the cytolytic response of NK cells. However, tumors have developed sophisticated strategies to evade NK cell surveillance, which lead to failure of tumor immunotherapy. In this paper, we summarized these immune escaping strategies, including the downregulation of ligands for activating receptors, upregulation of ligands for inhibitory receptors, secretion of immunosuppressive compounds, and the development of apoptosis resistance. Then, we focus on recent advancements in NK cell immune therapies, which include engaging activating NK cell receptors, upregulating NKG2D ligand MICA/B expression, blocking inhibitory NK cell receptors, adoptive NK cell therapy, chimeric antigen receptor (CAR)‐engineered NK cells (CAR‐NK), and NKG2D CAR‐T cells, especially several vaccines targeting MICA/B. This review will inspire the research in NK cell biology in tumor and provide significant hope for improving cancer treatment outcomes by harnessing the potent cytotoxic activity of NK cells.
Article
Immuno‐stimulative effect of chemotherapy (ISECT) is recognized as a potential alternative to conventional immunotherapies, however, the clinical application is constrained by its inefficiency. Metronomic chemotherapy, though designed to overcome these limitations, offers inconsistent results, with effectiveness varying based on cancer types, stages, and patient‐specific factors. In parallel, a wealth of preclinical nanomaterials holds considerable promise for ISECT improvement by modulating the cancer‐immunity cycle. In the area of biomedical nanomaterials, current literature reviews mainly concentrate on a specific category of nanomaterials and nanotechnological perspectives, while two essential issues are still lacking, i.e., a comprehensive analysis addressing the causes for ISECT inefficiency and a thorough summary elaborating the nanomaterials for ISECT improvement. This review thus aims to fill these gaps and catalyze further development in this field. For the first time, this review comprehensively discusses the causes of ISECT inefficiency. It then meticulously categorizes six types of nanomaterials for improving ISECT. Subsequently, practical strategies are further proposed for addressing inefficient ISECT, along with a detailed discussion on exemplary nanomedicines. Finally, this review provides insights into the challenges and perspectives for improving chemo‐immunotherapy by innovations in nanomaterials.
Article
Breast cancer (BC) is one of the most common malignancies in women worldwide. Numerous studies in immuno‐oncology and successful trials of immunotherapy have demonstrated the causal role of the immune system in cancer pathogenesis. The interaction between the tumor and the immune system is known to have a dual nature. Despite cytotoxic lymphocyte activity against transformed cells, a tumor can escape immune surveillance and leverage chronic inflammation to maintain its own development. Research on antitumor immunity primarily focuses on the role of the tumor microenvironment, whereas the systemic immune response beyond the tumor site is described less thoroughly. Here, a comprehensive review of the formation of the immune profile in breast cancer patients is offered. The interplay between systemic and local immune reactions as self‐sustaining mechanism of tumor progression is described and the functional activity of the main cell populations related to innate and adaptive immunity is discussed. Additionally, the interaction between different functional levels of the immune system and their contribution to the development of the pro‐ or anti‐tumor immune response in BC is highlighted. The presented data can potentially inform the development of new immunotherapy strategies in the treatment of patients with BC.
Article
The surface density of the triggering receptors responsible for the natural killer (NK)-mediated cytotoxicity is crucial for the ability of INK cells to kill susceptible target cells. In this study, we show that transforming growth factor beta1 (TGFbeta1) down-regulates the surface expression of NKp30 and in part of NKG2D but not that of other triggering receptors such as NKp46. The TGFbeta1-mediated inhibition of NKp30 surface expression reflects gene regulation at the transcriptional level. NKp30 has been shown to represent the major receptor involved in the NK-mediated killing of dendritic cells. Accordingly, the TGFbeta1-dependent down-regulation of NKp30 expression profoundly inhibited the NK-mediated killing of dendritic cells. On the contrary, killing of different NK-susceptible tumor cell lines was variably affected, reflecting the differential usage of NKp30 and/or NKG2D in the lysis of such tumors. Our present data suggest a possible mechanism by which TGFbeta1producing dendritic cells may acquire resistance to the NK-mediated attack.
Article
Tumours can be recognised by CTL and NK cells. CTL recognition depends on expression of MHC Class I loaded with peptides from tumour antigens. In contrast, loss of MHC Class I results in NK activation. In our study a large set of samples from patients with primary operable invasive breast cancer was evaluated for the expression of MHC Class I heavy and light by immunohistochemical staining of 439 breast carcinomas in a tissue microarray. Forty-seven percent (206 of 439) of breast carcinomas were considered negative for HLA Class I heavy chain (HC10), whereas lack of anti-β2m-antibody staining was observed in 39% (167 of 424) of tumours, with only 3% of the β2m-negative tumours expressing detectable HLA Class I heavy chain. Correlation with patient outcome showed direct relationship between patient survival and HLA-negative phenotype (log rank = 0.004). A positive relationship was found between the intensity of expression of MHC Class I light and heavy chains expression and histological grade of invasive tumour (p < 0.001) and Nottingham Prognostic Index (p < 0.001). To investigate whether HLA Class I heavy and light chains expression had independent prognostic significance, Cox multivariate regression analysis, including the parameters of tumour size, lymph node stage, grade and intensity of HC10 and anti-β2m staining, was carried out. In our analysis, lymph node stage (p < 0.001), tumour grade (p = 0.005) and intensity of MHC Class I light and heavy chains expression were shown to be independent prognostic factors predictive of overall survival (p-values HC10 = 0.047 and β2m = 0.018).
Article
For nearly a decade it has been appreciated that critical steps in human natural killer (NK) cell development likely occur outside of the bone marrow and potentially necessitate distinct microenvironments within extramedullary tissues. The latter include the liver and gravid uterus as well as secondary lymphoid tissues such as tonsils and lymph nodes. For as yet unknown reasons these tissues are naturally enriched with NK cell developmental intermediates (NKDI) that span a maturation continuum starting from an oligopotent CD34+CD45RA+ hematopoietic precursor cell to a cytolytic mature NK cell. Indeed despite the detection of NKDI within the aforementioned tissues, relatively little is known about how, why, and when these tissues may be most suited to support NK cell maturation and how this process fits in with other components of the human immune system. With the discovery of other innate lymphoid subsets whose immunophenotypes overlap with those of NKDI, there is also need to revisit and potentially re-characterize the basic immunophenotypes of the stages of the human NK cell developmental pathway in vivo. In this review, we provide an overview of human NK cell development in secondary lymphoid tissues and discuss the many questions that remain to be answered in this exciting field.
Article
Lymphocytes were originally thought to form the basis of a `cancer immunosurveillance' process that protects immunocompetent hosts against primary tumour development, but this idea was largely abandoned when no differences in primary tumour development were found between athymic nude mice and syngeneic wild-type mice. However, subsequent observations that nude mice do not completely lack functional T cells and that two components of the immune system-IFNγ and perforin-help to prevent tumour formation in mice have led to renewed interest in a tumour-suppressor role for the immune response. Here we show that lymphocytes and IFNγ collaborate to protect against development of carcinogen-induced sarcomas and spontaneous epithelial carcinomas and also to select for tumour cells with reduced immunogenicity. The immune response thus functions as an effective extrinsic tumour-suppressor system. However, this process also leads to the immunoselection of tumour cells that are more capable of surviving in an immunocompetent host, which explains the apparent paradox of tumour formation in immunologically intact individuals.