ArticlePDF Available

EpCAM Signaling Promotes Tumor Progression and Protein Stability of PD-L1 through the EGFR Pathway

Authors:

Abstract and Figures

Although epithelial cell adhesion molecule (EpCAM) has previously been shown to promote tumor progression, the underlying mechanisms remain largely unknown. Here, we report that the EGF-like domain I within the extracellular domain of EpCAM (EpEX) binds EGFR, activating both AKT and MAPK signaling to inhibit forkhead transcription factor O3a (FOXO3a) function and stabilize PD-L1 protein, respectively. Treatment with the EpCAM neutralizing antibody, EpAb2-6, inhibited AKT and FOXO3a phosphorylation, increased FOXO3a nuclear translocation, and upregulated high temperature requirement A2 (HtrA2) expression to promote apoptosis while decreasing PD-L1 protein levels to enhance the cytotoxic activity of CD8+ T cells. In vivo, EpAb2-6 markedly extended survival in mouse metastasis and orthotopic models of human colorectal cancer. The combination of EpAb2-6 with atezolizumab, an anti-PD-L1 antibody, almost completely eliminated tumors. Moreover, the number of CD8+ T cells in combination-treated tumors was increased compared with atezolizumab alone. Our findings suggest a new combination strategy for cancer immunotherapy in patients with EpCAM-expressing tumors. Significance This study shows that treatment with an EpCAM neutralizing antibody promotes apoptosis while decreasing PD-L1 protein to enhance cytotoxic activity of CD8+ T cells.
EpEX prevents cell apoptosis via inhibition of FOXO3a in colon cancer cells. A, HCT116 cells were starved and then treated with 2.5 mg/mL EpEX-Fc for the indicated times. Total cell lysate was examined by Western blotting. B, After treating with AG1478 (5 mmol/L) or wortmannin (1 mmol/L) for 1 hour, cells were exposed to EpEX (2.5 mg/mL) for 1 hour. The phosphorylation of FOXO3a was examined by Western blotting. C, The level of FOXO3a in nuclear fractions of EpEX-treated cells was probed (top). Serum-starved HCT116 cells were treated with EpEX-Fc for 2 hours and the location of FOXO3a was assessed by immunofluorescence (bottom). D, HCT116 cells were starved, pretreated with 20 mg/mL EpAb2-6, and then treated with EpEX for 15 minutes. The phosphorylation of AKT and FOXO3a was analyzed by Western blotting. E, HCT116 cells were treated with 20 mg/mL EpAb2-6 for 6 hours, and b-catenin and FOXO3a were detected in the nuclear fraction of HCT116 cells. F, qPCR analysis of FOXO3a-related gene expression (BIM, p21, and FasL) in control IgG-or EpAb2-6-treated HCT116 cells. G and H, HCT116 cell lines were treated with EpAb2-6 (in vitro; G) and mice bearing HCT116 subcutaneous xenograft were treated with EpAb2-6 (in vivo; H). Arrows, cellular localization of FOXO3a in the nucleus. Quantification of the number of cells with nuclear FOXO3a and b-catenin is shown when treated with EpAb2-6 in vitro (G) and in vivo (H; mean AE SD, N ¼ 200-300 cells in each group; Ã , P < 0.05). Bar graphs show mean AE SEM. Ã , P < 0.05; ÃÃ , P < 0.01.
… 
EpCAM is correlated with PD-L1 expression. A, The tumor growth of shLuc or shEpCAM H441 cells in NSG mice with or without PBMC injection. B, Relative tumor size was evaluated, and weights are shown in the graph. C, Evaluation of CD8 þ T cells in tumor tissue. D and E, Western blotting (D) and qRT-PCR analysis (E) of PD-L1 in H460 and H441 cells. F, Flow cytometry analysis of PD-L1 and EpCAM in H460 and H441 cells. G, H460 and H441 cells were treated with 50 mmol/L cycloheximide (CHX) for the indicated intervals and analyzed by Western blotting. Protein expression over time is shown in the graph. H, The protein and mRNA levels of PD-L1 in EpCAM-knockdown H441 cells were analyzed by Western blotting and qRT-PCR, respectively. I, EpCAM-knockdown H441 cells were treated with 50 mmol/L cycloheximide for the indicated intervals, and PD-L1 expression was analyzed by Western blotting. Protein expression over time is shown in the graph. J, The expression of exogenous PD-L1 in EpCAM-knockdown H441 cells was analyzed by Western blotting. K, IHC staining for EpCAM and PD-L1 was performed in the lung cancer tissue array, and the positive correlation between EpCAM and PD-L1 expression in patients with lung cancer is shown. L, GSEA enrichment profile of T-cell activation and proliferation in the lung cancer specimen. Statistical differences were determined by two-tailed Student t test. N ¼ 6 independent experiments. Graphical data are shown as mean AE SEM. ÃÃ , P < 0.01; ÃÃÃ , P < 0.001.
… 
Content may be subject to copyright.
CANCER RESEARCH | TUMOR BIOLOGY AND IMMUNOLOGY
EpCAM Signaling Promotes Tumor Progression and
Protein Stability of PD-L1 through the EGFR Pathway
Hao-Nien Chen, Kang-Hao Liang, Jun-Kai Lai, Chun-Hsin Lan, Mei-Ying Liao, Shao-Hsi Hung,
Yi-Ting Chuang, Kai-Chi Chen, William Wei-Fu Tsuei, and Han-Chung Wu
ABSTRACT
Although epithelial cell adhesion molecule (EpCAM) has pre-
viously been shown to promote tumor progression, the underlying
mechanisms remain largely unknown. Here, we report that the
EGF-like domain I within the extracellular domain of EpCAM
(EpEX) binds EGFR, activating both AKT and MAPK signaling to
inhibit forkhead transcription factor O3a (FOXO3a) function and
stabilize PD-L1 protein, respectively. Treatment with the EpCAM
neutralizing antibody, EpAb2-6, inhibited AKT and FOXO3a
phosphorylation, increased FOXO3a nuclear translocation, and
upregulated high temperature requirement A2 (HtrA2) expression
to promote apoptosis while decreasing PD-L1 protein levels to
enhance the cytotoxic activity of CD8
þ
T cells. In vivo, EpAb2-6
markedly extended survival in mouse metastasis and orthotopic
models of human colorectal cancer. The combination of EpAb2-6
with atezolizumab, an anti-PD-L1 antibody, almost completely
eliminated tumors. Moreover, the number of CD8
þ
T cells in
combination-treated tumors was increased compared with atezo-
lizumab alone. Our ndings suggest a new combination strategy for
cancer immunotherapy in patients with EpCAM-expressing
tumors.
Signicance: This study shows that treatment with an EpCAM
neutralizing antibody promotes apoptosis while decreasing PD-L1
protein to enhance cytotoxic activity of CD8
þ
T cells.
Introduction
Epithelial cell adhesion molecule (EpCAM) is the most frequently
expressed tumor-associated antigen; it is overexpressed in the majority
of human adenocarcinomas and squamous cell carcinomas and is
associated with poor prognosis in patients (14). EpCAM is sequen-
tially processed by a disintegrin and metalloproteinase 17 (ADAM17),
also called TNFaconverting enzyme (TACE), and g-secretase (5, 6).
Processing by these enzymes, respectively, releases the extracellular
domain (EpEX) and intracellular domain (EpICD). After release,
EpICD interacts with four and a half LIM domains protein 2 (FHL2)
and b-catenin to form a complex that translocates to the nucleus and
interacts with Lef-1, which binds to DNA (5, 6). The EpICD complex
promotes tumorigenesis of tumor-initiating cells through the upre-
gulation of reprogramming genes and epithelialmesenchymal tran-
sition (EMT; ref. 7). In addition, the increased release of EpEX
enhances EpICD generation and upregulates the expression of repro-
gramming and EMT genes (5, 79). It was previously reported that
EpEX could directly bind to EGFR and stimulate EGFR phosphory-
lation and its downstream signaling pathway (5, 8, 10). Moreover,
EpEX-induced EGFR phosphorylation can activate ADAM17 and
g-secretase to further increase the shedding of EpEX and
EpICD (5, 8, 10). However, the domain of EpEX that binds and
activates EGFR has not been previously identied.
AKT is a primary mediator of EGFR signaling that promotes cell
survival partially by inactivating proapoptotic proteins. One such
proapoptotic protein that is inactivated by AKT phosphorylation is
forkhead transcription factor O3a (FOXO3a; ref. 11). FOXO3a is also
referred to as FKHRL-1 and is a member of the forkhead transcription
factor family (12). Because genes activated by FOXO proteins generally
function to limit cell growth and promote death, this family is thought
of as tumor suppressors. When activated, FOXO3a accumulates in the
nucleus, where it enhances the transcription of various genes involved
in apoptosis and the cell-cycle control, such as BIM, FasL, and p21 (13).
Furthermore, AKT inactivation is an essential step in anoikis, and
FOXO3a regulation by the PI3K/AKT pathway may be essential for
this inactivation (14, 15). Thus, AKT inactivation has been suggested
to be an essential step in apoptosis (11).
In addition to its mediation of survival signals, EGFR activation is
crucial for triggering immune escape (16). EGFR-activating mutations
were reported to be associated with increased PD-L1 expression in
mouse models of EGFR-driven lung cancer and bronchial epithelial
cells with the expression of mutant EGFR (16). Moreover, EGFR
inhibitors can reduce PD-L1 expression in nonsmall cell lung cancer
(NSCLC) cell lines with activated EGFR, suggesting that EGFR sig-
naling may trigger immune escape (16). Further investigations dem-
onstrated that EGFR ligands, such as EGF, induce PD-L1 expression
primarily at the level of posttranslational modications. EGF was
shown to stabilize PD-L1 by inducing PD-L1 glycosylation, which
prevents GSK3-dependent proteasomal degradation of PD-L1 by
b-TrCP (17). Moreover, the MAPK signaling pathway, another impor-
tant axis of EGFR signaling, is associated with PD-L1 mRNA expres-
sion. It was reported that EGFR activation increases PD-L1 expression
through p-ERK1/2/p-c-Jun (1820). In addition to promoting PD-L1
transcription, MAPK signaling also stabilizes PD-L1 mRNA by atten-
uating TPP activity (21). Because EpCAM is an activator of EGFR
signaling, EpCAM might promote escape from immune surveillance.
However, EpCAM-mediated immunosuppression and its mechan-
isms are almost completely undescribed.
Institute of Cellular and Organismic Biology, Academia Sinica, Nankang, Taipei,
Taiwan.
Note: Supplementary data for this article are available at Cancer Research
Online (http://cancerres.aacrjournals.org/).
H.-N. Chen and K.-H. Liang contributed equally to this article.
Corresponding Author: Han-Chung Wu, Academia Sinica, 128 Academia Road,
Section 2, Nankang, Taipei 11529, Taiwan. Phone: 8862-2789-9528; Fax: 8862-
2788-0268; E-mail: hcw0928@gate.sinica.edu.tw
Cancer Res 2020;80:503550
doi: 10.1158/0008-5472.CAN-20-1264
2020 American Association for Cancer Research.
AACRJournals.org | 5035
Downloaded from http://aacrjournals.org/cancerres/article-pdf/80/22/5035/2796333/5035.pdf by MD ANDERSON HOSPITAL user on 14 June 2022
In this study, we report that the EGF-like domain I within EpEX
binds to EGFR and activates both the AKT and ERK1/2 pathways.
EpEX-induced AKT activation inhibits FOXO3a activity, which
decreases high temperature requirement A2 (HtrA2) transcription.
Moreover, a therapeutic antibody directed against EpCAM (EpAb2-6)
promotes FOXO3a nuclear translocation, and increases HtrA2 expres-
sion. We also found that EpCAM inhibits antitumor immunity by
activating the PD-1/PD-L1 pathway to suppress T-cell function. EpEX
increases PD-L1 protein stability rather than mRNA level, while
blockade of EGFR or MAPK signaling pathways attenuates EpEX-
mediated stabilization of PD-L1 protein. Furthermore, treatment of
EpAb2-6 markedly prolongs the survival of mice in metastasis and in
orthotopic models of human colorectal cancer, and a combination of
atezolizumab and EpAb2-6 shows enhanced therapeutic efcacy and
high levels of tumor-localized CD8
þ
T cells in a peripheral blood
mononuclear cell (PBMC) cell linederived xenograft (CDX) mouse
model. Our ndings not only shed light on the molecular mechanisms
underlying EpCAM signaling in cancer malignancy, but also suggest
that therapeutic targeting of EpCAM may work well in combination
with current immunotherapies.
Materials and Methods
Chemicals and antibodies
Antibodies against human EpCAM and p84 were purchased from
Abcam, and antibody against b-catenin was from Santa Cruz Bio-
technology. Anti-a-tubulin antibody was from Sigma-Aldrich. Poly-
clonal antibodies detecting total ERK and Thr202/Tyr204-phosphor-
ylated ERK, total AKT and Ser473-phosphorylated AKT, total
FOXO3a, Ser253-phosphorylated FOXO3a, Ser318/321-phosphory-
lated FOXO3a, and The32-phosphorylated FOXO3a, and active (non-
phospho Ser33/Ser37/Thr41) b-catenin, HtrA2, cytochrome c, COX
IV, XIAP, and PD-L1 were from Cell Signaling Technology. U0126
(MEK inhibitor), wortmannin (PI3K inhibitor), and MG132 (protea-
some inhibitor) were obtained from Selleck Chemicals.
Cell culture
Human lung adenocarcinoma cells (H441), lung carcinoma cells
(H460), embryonic kidney cells (HEK293T), colorectal carcinoma
cells (HCT116), and colorectal adenocarcinoma cells (SW620) were
obtained from the ATCC. H441 and H460 cells were cultured in
RPMI1640 Medium (Gibco), and HEK293T, HCT116, and SW620
cells were cultured in DMEM (Gibco). All cells were maintained
in conditioned medium supplemented with 10% FBS (Gibco) and
100 mg/mL penicillinstreptomycin (Gibco) at 37C in a humidied
incubator with 5% CO
2
.
Western blotting
Cells were lysed with RIPA buffer (20 mmol/L Tris-HCl, pH 7.4,
150 mmol/L NaCl, 1 mmol/L EDTA, 1 mmol/L EGTA, 1% Nonidet
P-40, 1% sodium deoxycholate, and 2.5 mmol/L sodium pyrophos-
phate) containing Protease Inhibitor (Roche) and Phosphatase Inhib-
itor (Roche) to extract whole-cell lysates. The lysates were then
quantied using the Pierce BCA Protein Assay Kit (Thermo Fisher
Scientic). The lysates were mixed with 5sample buffer (50 mmol/L
Tris-HCl, pH 6.8, 2% SDS, b-mercaptoethanol, 0.1% bromophenol
blue, and 10% glycerol), separated by SDS-PAGE, and transferred to
polyvinylidene diuoride (PVDF) Membrane (Millipore). Nonspecic
antibody-binding sites on the PVDF membrane were blocked with 3%
BSA in TBST (TBS buffer containing 0.1% Tween 20) and membranes
were incubated with indicated antibody overnight at 4C, followed by
incubation with horseradish peroxidaseconjugated secondary anti-
bodies (Jackson ImmunoResearch Laboratories) at room temperature
for 1 hour. The protein bands were subsequently visualized with
Chemiluminescence Reagents (Millipore) and detected by UVP
BioSpectrum 600 Imagining System (UVP). The protein expression
was quantied by Gel-Pro Analyzer 3.1 (Media Cybernetics).
Production and purication of EpEX-6xHis recombinant protein
The DNA fragment encoding EpEX (amino acids 24262 of
EpCAM) was amplied by PCR using PfuTurbo DNA polymerase
and cloned into pSecTag2 vector with C-terminal 6xHis tag to generate
pSecTag2-EpEX-6xHis. The EpEX-6xHis fusion protein was produced
with Expi293F Expression System (Thermo Fisher Scientic) and
puried by Ni-Afnity Column (GE Healthcare).
Extracellular interactions between EpEX-Fc and EGFR
HCT116 cells were harvested with 10 mmol/L EDTA in PBS, and
incubated with EpEX-Fc for 1 hour at 4C. After incubation, 2 mmol/L
DTSSP (Thermo Fisher Scientic) was used as a cross-linker to
stabilize the interaction between EpEX-Fc and EGFR. To stop the
cross-linking reaction, Tris, pH 7.5, was added to a nal concentration
of 20 mmol/L. Membrane proteins were extracted using Mem-PER
Eukaryotic Membrane Protein Extraction Reagent Kit (Thermo Fisher
Scientic). Finally, the EpEX-FcEGFR complex was pulled down with
Dynabeads Protein G (Invitrogen), and probed by Western blotting.
Construction of the EpCAM EGF-like domain deletion
In its EpEX, EpCAM contains two EGF-like domains, at amino
acids 2759 (rst EGF-like domain) and 66135 (second EGF-like
domain), and a cysteine-free motif (22). The EpCAM EGF-like
domain deletion was generated using the standard QuikChange
deletion mutation system with rst forward mutagenic deletion primer
(50-GCAGCTCAGGAAGAATCAAAGCTGGCTGCC-30) and rst
reverse mutagenic deletion primer (50-GGCAGCCAGCTTTGATTC-
TTCCTGAGCTGC-30); and second forward primer (50-AAGCTGG-
CTGCCAAATCTGAGCGAGTGAGA-30) and second reverse primer
(50-TCTCACTCGCTCAGATTTGGCAGCCAGCTT-30). The PCR
amplications were performed using KAPA HiFi Hot Start DNA
Polymerase (Kapa Biosystems), and products were treated with restric-
tion enzyme, DpnI (Thermo Fisher Scientic), to digest methylated
parental DNAs.
Cycloheximide chase assay
Cycloheximide, a protein synthesis inhibitor, was used to evaluate
the stability of PD-L1. Cells were treated with cycloheximide for 0, 2, 4,
or 6 hours. Proteins were extracted and Western blot analysis was
performed to detect PD-L1 protein level.
Immunoprecipitation assay
Cells were lysed in lysis buffer (50 mmol/L Tris-HCl, pH 7.4, 150
mmol/L NaCl, and 1% NP-40) with Protease Inhibitors (Roche). For
immunoprecipitation, cell lysates were incubated with antibodies for
6 hours at 4C. Then, 20 mL Dynabeads Protein G was added and the
mixture was incubated for 2 hours at 4C to pull-down the antibody-
bound protein. The immunoprecipitation samples were washed with
PBS three times, denatured in sample buffer, and analyzed by Western
blotting.
Generation of mAbs and purication of IgG
Generation of EpAb2-6 and control IgG was performed as described
previously (23). The protocol was approved by the Committee on the
Chen et al.
Cancer Res; 80(22) November 15, 2020 CANCER RESEARCH
5036
Downloaded from http://aacrjournals.org/cancerres/article-pdf/80/22/5035/2796333/5035.pdf by MD ANDERSON HOSPITAL user on 14 June 2022
Ethics of Animal Experiments of Academia Sinica [Nankang, Taipei,
Taiwan, AS Institutional Animal Care and Use Committee (IACUC):
11-04-166].
Lentivirus-mediated short hairpin RNA knockdown
All lentiviral short hairpin RNA (shRNA) constructs were pur-
chased from the RNAi Core Facility at Academia Sinica (Nankang,
Taipei, Taiwan). Lentiviral production, infection, and selection were
performed according to protocols from the RNAi Consortium
(Academia Sinica, Nankang, Taipei, Taiwan). To produce the lenti-
virus, HEK293T cells were transiently cotransfected with shRNA
plasmid, packaging (pCMV-DR8.91) and envelope (pMD.G) expres-
sion plasmids, using PolyJet DNA Transfection Reagent (SignaGen
Laboratories). The next day, the transfection solution was replaced
with 1% BSA containing medium to improve virus yield. Two days
after transfection, medium containing lentivirus was collected. Cells
were cultured in lentivirus-containing medium, supplemented with
8mg/mL polybrene, for another 48 hours. The transduced cells were
selected with 2 mg/mL puromycin for 4 days and the knockdown
efciency was measured by Western blotting. The target sequence for
human EpCAMspecic shRNA was shRNA 50-GCAAATGGACA-
CAAATTAC AA-30. Luciferase shRNA (shLuc) was used as a negative
control.
RNA extraction, cDNA synthesis, quantitative reverse-
transcription PCR
Total RNA extraction, rst-strand cDNA synthesis, and SYBR
Greenbased real-time PCR were performed as described in the
manufacturers instructions. To extract total RNA, cells were lysed
using TRIzol Reagent (Invitrogen), and proteins and phenol were
removed from TRIzol using chloroform. After centrifugation, the
top colorless layer was collected and mixed with isopropanol to
precipitate RNA pellet. Then, the RNA pellet was washed with
70% ethanol, air-dried at room temperature, and dissolved in
RNase-free water. For rst-strand cDNA synthesis, 5 mgoftotal
RNA was used for reverse transcription with oligo(dT) primer
and SuperScriptIII Reverse Transcriptase (Invitrogen) at 50Cfor
60 minutes. Target gene levels were evaluated by quantitative PCR
(qPCR), using LightCycler 480 SYBR Green I Master Mix (Roche)
and a LightCycler480 System (Roche). GAPDH mRNA expression
was measured as endogenous housekeeping control to normalize
all qPCR reactions. The qPCR reaction was as follows: 95Cfor
5 minutes, followed by 40 cycles of denaturation at 95Cfor
10 seconds, annealing at 60C for 10 seconds, and extension at
72C for 30 seconds. Final results were calculated from three
independent experiments. Primer sequences used to detect the
mRNA expression of genes of interest are listed in Supplementary
Table S1.
Immunouorescence assay
For immunouorescence, 2 10
4
cells were seeded on 12-mm
cover slips in 24-well culture plate, xed with 2% paraformalde-
hyde, and incubated with 0.1% Triton X-100 to increase the
permeability of cell membrane. Then, samples were blocked with
1% BSA for 1 hour at room temperature and stained with
primary antibodies at 4C overnight. Slides were washed and
stained with FITC or Alexa568-conjugated secondary antibodies
and DAPI for 40 minutes. Cover slips were then washed with
PBS and mounted in Mounting Solution (Vector Laboratories).
The slides were examined with a confocal microscope (Leica
TCS-SP5).
Apoptosis protein array
HCT116 cells were stimulated with 20 mg/mL EpAb2-6 for 6 hours,
and apoptosis arrays were performed according to the manufacturers
instructions (R&D System; ARY009). The membranes (arrays) were
detected by the UVP BioSpectrum 600 Imagining System (UVP). The
arrays were quantied by a Gel-Pro Analyzer 3.1 (Media Cybernetics,
Inc.).
Flow cytometry analysis
Cells were harvested, washed, and suspended in FACS buffer (PBS
with 1% FBS), and 10
5
cells were transferred to 96-well shrill-based
plates. Cells were stained with different antibodies for 1 hour at 4C,
and then subsequently stained with indicated PE-conjugated second-
ary antibodies for 1 hour at 4C. Then, cells were washed with FACS
buffer twice and suspended in 400 mL FACS buffer. Florescence signals
were analyzed using ow cytometry (BD FASCSC AntoII) and mea-
sured with FCS Express V3 software. The data were collected from
three independent experiments.
Chromatin immunoprecipitation
In brief, HCT116 cells (1 10
5
) were treated with 20 mg/mL EpAb2-
6 for 6 hours, followed by cross-linking xation in 1% formaldehyde.
Fixation was quenched by the addition of glycine to a nal concen-
tration of 200 mmol/L and xed chromatin complexes were then
sonicated to an average length of 250 bp using an MISONIX Sonicator
3000. The sonicated proteinDNA complexes were subjected to
immunoprecipitation using 2 mg of antibodies against FOXO3a. The
immunoprecipitated DNA was recovered by a PCR Purication Kit
(Qiagen), and the amount of target DNA was detected by PCR using
the primers complementary to the sequence listed in Supplementary
Table S2.
Dual-luciferase reporter assays
The human HtrA2 proximal promoter fragments covering the
regions 1628 to þ86 and 700 to þ86 (with the transcriptional
start site denoted as þ1) were PCR amplied and then cloned into the
rey luciferase reporter plasmid pGL4.18 Vector (Promega). To
assess the effect of EpA2-6 on HtrA2 promoter activity, HCT116 and
SW620 cells (1 10
4
/well), seeded onto 24-well dishes, were tran-
siently transfected with HtrA2 promoter reporter constructs in com-
bination with a plasmid expressing Renilla luciferase using the PolyJet
Transfection Reagent (SignaGen Laboratories) for 24 hours, followed
by 20 mg/mL EpAb2-6 for 6-hour stimulation. Cell lysates were
prepared and subjected to the luciferase activity assay using the
Dual-Luciferase Reporter Assay Kit (Promega). Firey luciferase
activity was normalized to Renilla luciferase activity, and nal data
were presented as the fold induction of luciferase activity compared
with EpAb2-6-IgG controls. Data are expressed as means SD from
three independent experiments. Primers used to amplify the human
HtrA2 proximal promoter fragment are listed in Supplementary
Table S3.
Apoptosis and mitochondrial membrane potential assay
Cells (2 10
5
) were seeded onto 24-well dishes and treated with
20 mg/mL control IgG or EpAb2-6 for 6 hours. The apoptotic cell and
mitochondrial membrane potentials were detected using FITC
Annexin V Apoptosis Detection Kit (BD Biosciences) and MitoStatus
Red (BD Biosciences), respectively. The apoptotic cell and mitochon-
drial membrane potentials were analyzed using a Flow Cytometer
(Thermo Fisher Scientic). Each measurement was carried out
at least three times to ensure reproducibility. The effect of gene
EpCAM Signaling Regulates Tumorigenesis
AACRJournals.org Cancer Res; 80(22) November 15, 2020 5037
Downloaded from http://aacrjournals.org/cancerres/article-pdf/80/22/5035/2796333/5035.pdf by MD ANDERSON HOSPITAL user on 14 June 2022
knockdown on EpAb2-6induced apoptosis is represented in Supple-
mentary Table S4.
IHC
Human lung cancer tissue microarray was purchased from Super
BioChip. After exhausting the endogenous hydroperoxidases with 3%
hydrogen peroxide in methanol for 30 minutes, the tissue microarray
was blocked with 1% BSA for 1 hour and exposed to anti-PD-L1 (Clone
28-2, Abcam) or EpAb3-5 (developed by our laboratory as described
previously; ref. 23) at 4C overnight. After washing with PBST, the
tissue microarray was processed using the standard procedure of Super
Sensitive IHC Detection System (Bio-Genex) and stained by DAB.
Nuclei were stained with Mayers hematoxylin solution (Wako).
PBMC and T-cell isolation and culture
For the experiments using PBMCs, blood samples were drawn from
healthy donors and collected into 10 mL Vacutainer tubes containing
the anticoagulant EDTA (BD Bioscience). After centrifugation at
1,500 rpm for 10 minutes, plasma was removed from the sample and
mixed with an equal volume of PBS. The mixture was layered on Ficoll-
Paque plus (Ficoll:blood ¼1:2) and centrifuged at 1,500 rpm for 30
minutes. The PBMC layer (buffy coat) was collected and washed with
PBS with 0.5% BSA and 2 mmol/L EDTA twice. To obtain puried
CD3
þ
T cells, PBMCs were positively selected by anti-CD3 magnetic
beads (MACS), according to the standard protocol. Isolated CD3
þ
T cells (1 10
6
) were cultured and activated by 25 mL anti-CD3/
anti-CD28coated Dynabeads (Invitrogen) in RPMI1640 medium
supplemented with 10% FBS, 100 mg/mL penicillinstreptomycin,
12.5 ng/mL IL2 (Gibco), and 1 ng/mL IL15 (MACS) for 48 hours.
The protocol was approved by the Institutional Review Board (IRB)
of Academia Sinica (Nankang, Taipei, Taiwan, AS IRB: AS-IRB01-
19049).
Cell viability assays in vivo
NOD/SCID mice were intravenously injected with HCT116-GFP
cells; mice bearing circulating HCT116-GFP cells were intravenously
treated with EpAb2-6 or an equivalent dosage of control IgG at 1 hour
after cell injection (antibody was delivered at 20 mg/kg). Then, blood
samples were obtained from the facial vein of mice and the uores-
cence intensity of whole blood was quantied at the indicated time-
points. The uorescence was measured with a Microplate Reader
(Molecular Devices, SpectraMax M5) at an excitation wavelength of
355 nm and emission wavelength of 440 nm.
Subcutaneous human colorectal cancer studies
Subcutaneous studies were performed as reported previously (23).
The protocol was approved by the Committee on the Ethics of Animal
Experiments of Academia Sinica (Nankang, Taipei, Taiwan, AS
IACUC: 11-04-166).
Orthotopic implantation and therapeutic studies
Orthotopic studies were performed as reported previously (24).
Briey, NSG mice (NOD.Cg-Prkdcscid Il2rgtm1Wjl/SzJ) were used
for orthotopic implantation of HCT116 cells, which were previously
infected with Lenti-luc virus (lentivirus containing luciferase genes).
The mice were anesthetized by intraperitoneal injection of avertin,
2,2,2-Tribromo-Ethanol (Sigma-Aldrich) at a dose of 250 mg/kg.
Tumor development was monitored by bioluminescence imaging. For
the orthotopic therapeutic study, tumor-bearing mice were treated
with control IgG or EpAb2-6 (20 mg/kg). Tumor progression was
monitored by quantication of bioluminescence. Mouse body weight
and survival rate were measured. Animal care was carried out in
accordance with the guidelines of Academia Sinica (Nankang, Taipei,
Taiwan). The protocol was approved by the Committee on the Ethics
of Animal Experiments of Academia Sinica (Nankang, Taipei, Taiwan,
AS IACUC: 11-04-166).
Gene set enrichment analysis
The gene expression matrix was obtained from LUAD projects in
The Cancer Genome Atlas. The 25% of samples with highest EpCAM
expression were dened as the EpCAM_Highgroup and the
25% samples with least EpCAM expression were dened as the
EpCAM_Lowgroup. T-cell activation and proliferation gene sets
provided by gene set enrichment analysis (GSEA) website were then
used to analyze the data.
Statistical analyses
All data are represented as means SD from at least three
independent experiments. Signicant differences from the respective
control for each experimental condition were calculated using
Student ttest, unless otherwise specied. ,P<0.05; ,P<0.01; or
,P<0.001 are indicated as signicant. Survival analysis was
performed using a log-rank test. The correlation coefcient was
calculated using Spearman analysis.
Results
EpEX binds to EGFR through its EGF-like domain I
Previously, we reported that EpEX can bind to EGFR (10), and
functions as a growth factor and induces EGFR phosphorylation (5, 10).
To verify that EpEX interacts with EGFR, we used the cross-linker,
DTSSP, to stabilize the EpEXEGFR complex. The binding of EpEX to
EGFR was conrmed by immunoprecipitation and Western blotting
(Fig. 1A). To evaluate the direct binding of recombinant EpEX to the
EpEX of EGFR (EGFR
ECD
), we performed ELISA to probe the direct
interaction between puried EpEX and EGFR
ECD
protein (Fig. 1B).
To test whether membrane-bound EpCAM could bind to EGFR, we
performed immunoprecipitation experiments using HEK293T cells
with EpCAM-V5 and EGFR-Flag overexpression. The interaction
between exogenous EpCAM and EGFR was detected by coimmuno-
precipitation (Fig. 1C). In EpCAM EpEX, it contains two EGF-like
domains, at amino acids 2759 (rst EGF-like domain) and 66135
(second EGF-like domain), and a cysteine-free motif (22). To identify
the specic region of EpEX that binds to EGFR, we constructed
different EGF-like domaindeleted EpCAM mutants. Surprisingly,
the EGF-like domain Ideleted EpCAM mutant (EpCAM
DEGFI
)
showed decreased binding to EGFR, while the EGF-like domain II
deleted EpCAM mutant (EpCAM
DEGFII
) exhibited an enhanced inter-
action with EGFR (Fig. 1D). To further evaluate the binding of soluble
EpEX to EGFR
ECD
, HEK293T cells were cotransfected with different
vectors expressing soluble EGFR
ECD
-Flag or EpEX-Fc, and the protein
complex was examined in culture media (Fig. 1E). Soluble EGF-like
domain IIdeleted EpEX-Fc (EpEX
DEGFII
-Fc) showed signicantly
increased afnity, while EGF-like domain Ideleted EpEX-Fc
(EpEX
DEGFI
-Fc) exhibited decreased afnity for EGFR
ECD
-Flag com-
pared with EpEX-Fc (Fig. 1F). Overall, we found that the membrane-
bound EpCAM and secreted EpEX both bind EGFR and deletion of the
EGF domain I diminished this binding (Fig. 1CF).
Similar results were observed using puried wild-type or mutant
EpEX and EGFR
ECD
recombinant proteins. The recombinant
EpEX
DEGFII
protein had a stronger binding afnity for EGFR
ECD
than
wild-type controls, and the EpEX
DEGFI
protein lost its ability to bind
Chen et al.
Cancer Res; 80(22) November 15, 2020 CANCER RESEARCH
5038
Downloaded from http://aacrjournals.org/cancerres/article-pdf/80/22/5035/2796333/5035.pdf by MD ANDERSON HOSPITAL user on 14 June 2022
Figure 1.
EpEX bindsto EGFR through its EGF-likedomain I. A, Immunoprecipitation (IP)of afnity cross-linkedEpEX-Fc bound to EGFR from HCT116cells. B, EpEX-Fc was added
to EGFR
ECD
-coated ELISA plates and detected by TMB (3,30,5,50-tetramethylbenzidine) colorimetric peroxidase assay. C, 293T cells were transfected with EGFR-Flag
and EpCAM-V5. Immunoprecipitation was performed with anti-V5 antibody (left) or anti-Flag antibody (right), followed by Western blotting. D, 293T cells were
transfected with EGFR-Flag and EGF domaindeleted mutant EpCAM-V5. The protein interaction was probed by immunoprecipitation with anti-V5 antibody and
Western blotting with anti-Flag antibody. E, In culture media from 293T cells with EGFR
ECD
-Flag and EpEX-Fc transfection, immunoprecipitation was performed with
DynabeadsProtein G and detected byWestern blotting withanti-Flag antibody.F, 293T cells were transientlytransfected with EGFR
ECD
-Flag and EGFdomaindeleted
mutant EpEX-Fc. The EGFR
ECD
-Flag and mutant EpEX-Fc interaction was examined by immunoprecipitation with Dynabeads Protein G and Western blotting with
anti-Flag antibody. G, Puried EGFR
ECD
-6xHis and EGF domaindeleted mutant EpEX-Fc were incubated and immunoprecipitation was performed with Dynabeads
Protein G and detected by Western blotting. H, HCT116 cells were starved and treated with wild-type or EGF domaindeleted mutant EpEX, and EGFR signaling was
analyzed by Western blotting.N¼3 independent experiments. Graphical data are shown as meanSEM. ,P<0.001, two-tailed Studentttest. IB, immunoblotting.
EpCAM Signaling Regulates Tumorigenesis
AACRJournals.org Cancer Res; 80(22) November 15, 2020 5039
Downloaded from http://aacrjournals.org/cancerres/article-pdf/80/22/5035/2796333/5035.pdf by MD ANDERSON HOSPITAL user on 14 June 2022
EGFR
ECD
(Fig. 1G). Moreover, the EpEX
DEGFII
protein induced EGFR
signaling like the wild-type control, but EpEX
DEGFI
protein did not
(Fig. 1H). These results suggest that the EGF-like domain I in EpEX is
the major domain responsible for the EpEXEGFR interaction and
activation of EGFR signaling.
Inhibiting EpEX increases FOXO3a nuclear accumulation and
induces apoptosis
Because we found that EpEX activates EGFR through EpEX domain
I, we further investigated whether EpEX-mediated EGFR activation is
important for cancer progression. AKT is a major downstream effector
of EGFR, which promotes cell survival partially by inactivating
FOXO3a (12). Inhibition of AKT/FOXO3a signaling has been sug-
gested an essential step in apoptosis (11) and could inhibit tumor-
sphere formation of the SKOV3 ovarian cancer cell line (25). To
evaluate whether EpEX affects FOXO3a function, we examined
FOXO3a phosphorylation and intracellular location. Immunoblotting
analysis demonstrated that EpEX induced the phosphorylation of AKT
and FOXO3a in a time-dependent manner (Fig. 2A), and EpEX-
induced phosphorylation of FOXO3a was inhibited by AG1478 (EGFR
inhibitor) or wortmannin (PI3K inhibitor; Fig. 2B). Moreover, after
EpEX treatment, FOXO3a translocated from the nucleus to the
cytoplasm (Fig. 2C). Because the EpEX acted through EGFR, we used
EGF, a well-known ligand for EGFR, as an inducer. Similar to EpEX
treatment, EGF induced nuclear exclusion of FOXO3a (Supplemen-
tary Fig. S1AS1C). Therefore, we conclude that EpEX inhibits the
FOXO3a activity through the EGFRAKT pathway.
Previously, we had developed a neutralizing antibody, EpAb2-6,
which targets EpEX and induces apoptosis (5, 23). EpAb2-6 was
employed to block the EpEX. By blocking the function of EpEX,
EpAb2-6 treatment is known to interrupt the EpEX/EGFR/ADAM17
axis, which represents a positive feedback loop promoting EpCAM
cleavage and subsequently increases EpEX and EpICD production (5).
Indeed, decreases in ADAM17 and g-secretase activity and soluble
EpEX release were observed after EpAb2-6 treatment (Supplementary
Fig. S2AS2C). Moreover, nuclear-activated b-catenin and down-
stream genes, including reprogramming genes and EMT-related
genes, were also decreased in EpAb2-6treated cells (Supplementary
Fig. S2D and S2E). Next, we examined whether EpCAM contributes to
anchorage-independent growth. For this purpose, a soft agar colony
formation assay was performed with colon cancer cell lines. EpAb2-6
treated cells formed signicantly smaller and fewer colonies compared
with the control cells (Supplementary Fig. S2F). We further cultivated
these attached colon cancer cells into tumorspheres and found that
EpAb2-6 treatment inhibited tumorsphere formation (Supplementary
Fig. S2G). These results suggest that targeting EpCAM with a specic
antibody, EpAb2-6, is a potentially viable strategy to suppress colon
cancer malignancy.
It has been previously reported that the nuclear accumulation of
FOXO3a promotes cancer cell apoptosis after treatment with various
chemotherapy drugs or EGFR inhibitors (2628). Thus, we wanted to
clarify whether inhibiting the function of EpEX facilitates nuclear
accumulation of FOXO3a. Blocking the function of EpEX with EpAb2-
6 treatment also decreased the phosphorylation of AKT and FOXO3a
(Fig. 2D) and subsequently increased nuclear accumulation of
FOXO3a, while diminishing the nuclear translocation of b-catenin
(Fig. 2E). The downstream genes of FOXO3a (BIM, p21, and FASL)
were also upregulated in HCT116 cells with EpAb2-6 treatment
compared with IgG control (Fig. 2F). Using immunouorescence
staining, FOXO3a and b-catenin were detected in both the cytoplasm
and nucleus of colon cancer cells. However, after EpAb2-6 treatment,
FOXO3a was increased in the nucleus, but nuclear b-catenin was
decreased both in vitro (Fig. 2G) and in vivo (Fig. 2H). Thus,
inhibiting EpEX enhances the nuclear translocation of FOXO3a and
consequently upregulates downstream genes, which are involved in
promoting apoptosis.
To evaluate whether EpAb2-6 induces apoptosis via EGFR/AKT/
FOXO3a, we used shRNA to knockdown EGFR, AKT, and FOXO3a
expression in colon cancer cells and examined the effect of EpAb2-6 on
cancer cell apoptosis. EGFR, AKT, and FOXO3a-knockdown cells
were all less sensitive to EpAb2-6induced apoptosis than control cells,
and EpAb2-6induced apoptosis was completely abrogated in
EpCAM-knockout cells (Supplementary Fig. S3AS3D). Together,
our data show that EpEX induces phosphorylation of FOXO3a and
inactivates its biological function; meanwhile, blocking the function of
EpEX inhibits the phosphorylation of FOXO3a to promote nuclear
accumulation of FOXO3a and activates expression of its downstream
proapoptotic genes.
HtrA2 acts downstream of FOXO3a and contributes to
EpAb2-6induced apoptosis
To gain further insight into the mechanisms of EpAb2-6induced
apoptosis, we compared the expression of 35 apoptosis-related sig-
naling proteins in control IgG- and EpAb2-6treated colon cancer
cells. We found that among the 35 apoptosis-related proteins, two
showed substantial increases in expression after EpAb2-6 treatment
compared with the IgG control. These proteins were HtrA2 and
X-linked inhibitor of apoptosis protein (XIAP; Fig. 3A).
Immunoblotting and qPCR experiments showed that EpAb2-6
treatment enhanced the protein and mRNA expression of HtrA2 in
HCT116 cells (Fig. 3B). According to previous studies, HtrA2 can be
released into the cytosol, where it contributes to apoptosis (29). As
expected, we found that EpAb2-6 induced the release of cytochrome c
and HtrA2 from mitochondria into the cytosol (Fig. 3C). Further-
more, EpAb2-6 treatment induced the dissipation of the mitochon-
drial membrane potential (Fig. 3D).
HtrA2 mRNA levels were increased after EpAb2-6 treatment,
implying transcriptional regulation of HtrA2 gene was induced by
EpAb2-6. To examine this hypothesis, HCT116 cells were transiently
transfected with a reporter for the HtrA2 promoter; a region encom-
passing 1,704 bases upstream of the HtrA2 translational start site was
used to drive the expression of the rey luciferase gene (pGL4.18-
HtrA2-1;Fig. 3E). EpAb2-6 treatment led to an approximate 4-fold
induction in HtrA2 promoter activity, compared with the IgG control,
indicating that the HtrA2 gene is indeed a transcriptional target of
EpAb2-6induced signaling (Fig. 3E). To further elucidate the molec-
ular mechanism controlling HtrA2 gene transcription, the HtrA2
promoter sequence was analyzed using the PROMO virtual laboratory
website. This analysis uncovered putative cis-acting response elements
(1283 to 1299) for FOXO transcription factors. Interestingly, to
our knowledge, there is no previously published evidence showing that
FOXO3a controls HtrA2 promoter activity. Thus, we examined
whether HtrA2 promoter activity was regulated by FOXO3a by
generating a reporter construct without the FOXO3a response element
(pGL4.18-HtrA2-2), as depicted in Fig. 3E. Indeed, the EpAb2-6
induced increase in luciferase activity was abolished in the absence of a
FOXO3a response element (Fig. 3E).
To further elucidate whether FOXO3a directly regulates the HtrA2
promoter, chromatin immunoprecipitation (ChIP) analysis was per-
formed using primers anking the putative FOXO3a response ele-
ments (1283 to 1299). FOXO3a occupied the HtrA2 promoter after
EpAb2-6 treatment, which was not observed in FOXO3a-knockdown
Chen et al.
Cancer Res; 80(22) November 15, 2020 CANCER RESEARCH
5040
Downloaded from http://aacrjournals.org/cancerres/article-pdf/80/22/5035/2796333/5035.pdf by MD ANDERSON HOSPITAL user on 14 June 2022
Figure 2.
EpEX prevents cell apoptosis via inhibition of FOXO3a in colon cancer cells. A, HCT116 cells were starved and then treated with 2.5 mg/mL EpEX-Fc for the indicated
times. Total cell lysate was examined by Western blotting. B, After treating with AG1478 (5 mmol/L) or wortmannin (1 mmol/L) for 1 hour, cells were exposed to EpEX
(2.5 mg/mL) for 1 hour. The phosphorylation of FOXO3a was examined by Western blotting. C, The level of FOXO3a in nuclear fractions of EpEX-treated cells was
probed (top). Serum-starved HCT116 cells were treated with EpEX-Fc for 2 hours and the location of FOXO3a was assessed by immunouorescence (bottom).
D, HCT116 cells were starved, pretreated with 20 mg/mL EpAb2-6, and then treated with EpEX for 15 minutes. The phosphorylation of AKT and FOXO3 awas analyz ed
by Western blotting. E, HCT116 cells were treated with 20 mg/mL EpAb2-6 for 6 hours, and b-catenin and FOXO3a were detected in the nuclear fraction of HCT116
cells. F, qPCR analysis of FOXO3a-related gene expression (BIM, p21,andFasL) in control IgG- or EpAb2-6treated HCT116 cells. Gand H, HCT116 cell lines were
treated with EpAb2-6 (in vitro;G) and mice bearing HCT116 subcutaneous xenograft were treated with EpAb2-6 (in vivo;H). Arrows, cellular localization of FOXO3a
in the nucleus. Quantication of the number of cells with nuclear FOXO3a and b-catenin is shown when treated with EpAb2-6 in vitro (G)andin vivo (H; mean SD,
N¼200300 cells in each group; ,P<0.05). Bar graphs show mean SEM. ,P<0.05; ,P<0.01.
EpCAM Signaling Regulates Tumorigenesis
AACRJournals.org Cancer Res; 80(22) November 15, 2020 5041
Downloaded from http://aacrjournals.org/cancerres/article-pdf/80/22/5035/2796333/5035.pdf by MD ANDERSON HOSPITAL user on 14 June 2022
cells (Fig. 3F). These results indicate that FOXO3a increased HtrA2
transcription through direct binding to the HtrA2 promoter. We next
investigated whether HtrA2 plays an important role in EpAb2-6
induced cell death. In this experiment, we used shRNA to knockdown
HtrA2 expression in colon cancer cells. In HtrA2-knockdown cells,
EpAb2-6 treatment induced fewer apoptotic cells (Fig. 3G) and less
disruption of mitochondrial membrane potential (Fig. 3H) than shLuc
(control knockdown) cells. Hence, FOXO3a-mediated HtrA2 tran-
scription appears to be involved in EpAb2-6induced apoptosis.
EpCAM expression is positively correlated with PD-L1
stabilization
EGFR activation was reportedly associated with increased PD-L1
expression in mouse models of EGFR-driven lung cancer and bron-
chial epithelial cells with mutant EGFR expression, suggesting that
EGFR signaling is crucial for triggering immune escape (16). Further
investigations then demonstrated that increased EGFR signaling
upregulates PD-L1 expression by various mechanisms (17, 19, 21).
Moreover, EpCAM plays a crucial role in EGFR activation (5, 8, 10).
Figure 3.
EpEX prevents apoptosis by inhibiting FOXO3a-mediated HtrA2 expression. A, HCT116 cells were treated with control IgG or EpAb2-6 for 6 hours, and apoptosis-
related proteins were probed by the human apoptosis array kit. B, Expression of HtrA2 was examined by immunoblotting and qPCR analysis. C, EpAb2-6 induced the
release of mitochondrial apoptogenic proteins, HtrA2 and cytochrome c, in colon cancer cells. D, The percentage of cells with mitochondrial membrane potential
(Dcm) depolarization was measured after EpAb2-6 treatment. E, HCT116 cells with transfection of human HtrA2 proximal promoterdriven luciferas e reporter were
treated with control IgG or EpAb2-6. Luciferase activity assays were performed to evaluate the activity of the HtrA2 promoter. Results are expres sed as fold induction
compared with the IgG control. Statistical differences were determined by one-way ANOVA and Bonferroni multiple comparison test. F, Quantitative ChIP analysis of
FOXO3a binding to HtrA2 promoters was performed. G, HtrA2-knockdown HCT116 cells were treated with EpAb2-6, and the apoptotic cells were quantied by
uorescein Annexin V-FITC/propidium iodide double labeling. Statistical differences were determined by one-way ANOVA and Bonferroni multiple comparison test.
H, HtrA2-knockdown HCT116 cells were treated with EpAb2-6, and the percentage of cells with mitochondrial membrane potential (Dcm) depolarization was
measured. Statistical differences were determined by one-way ANOVA and Bonferr oni multiple comparison test. N¼3 independent experiments. Graphical data are
shown as mean SEM. ,P<0.05; ,P<0.01; ,P<0.001. TSS, transcription start site.
Chen et al.
Cancer Res; 80(22) November 15, 2020 CANCER RESEARCH
5042
Downloaded from http://aacrjournals.org/cancerres/article-pdf/80/22/5035/2796333/5035.pdf by MD ANDERSON HOSPITAL user on 14 June 2022
On the basis of these ndings, we decided to investigate whether
EpCAM regulates PD-L1 expression via EGFR activation. Because
HCT116 cells expression of PD-L1 is extremely low, H441 cells with
high expression of PD-L1 were substituted for HCT116 cells as an
experimental model for further studies.
First, we wanted to clarify whether EpCAM is involved in escape
from immune surveillance. Immunodecient NSG mice carrying
shLuc H441 xenografts in the left side and shEpCAM H441 xenografts
in the right side were intravenously injected with PBMCs to recon-
stitute immune cells. EpCAM-knockdown cells exhibited decreased
tumor weight, conrming that EpCAM promotes tumor progression
(Fig. 4A and B). Interestingly, shLuc tumors in mice with PBMC
injection were not different from shLuc tumors in mice without
PBMCs. However, shEpCAM tumors with PBMCs were smaller than
those without PBMCs, implying that EpCAM might inhibit antitumor
immunity (Fig. 4A and B). By ow cytometry analysis, we found that
CD8
þ
T cells were enriched in shEpCAM tumor tissue (Fig. 4C), and
Western blot analysis demonstrated that PD-L1 expression was
diminished in shEpCAM tumor tissue (Fig. 4D). These results indicate
that EpCAM promotes PD-L1 expression and reduces tumor-
associated CD8
þ
T cells in vivo.
To investigate the correlation between EpCAM and PD-L1 expres-
sion, we analyzed PD-L1 mRNA expression in two NSCLC cell lines,
H441 (high EpCAM expression) and H460 (no EpCAM expression).
Compared with H441 cells, H460 cells exhibited a higher level of PD-
L1 mRNA, but a lower level of PD-L1 protein (Fig. 4E). Flow
cytometry analysis of PD-L1 and EpCAM conrmed this nding
(Fig. 4F). Because many previous studies had indicated that the
protein stability of PD-L1 plays an important role in its
expression (17, 3032), and, on the basis of our results, we suspected
that stability of PD-L1 might serve as a key regulator of protein level.
Indeed, a cycloheximide chase assay indicated that the half-life of PD-
L1 in H460 cells was reduced compared with that in H441 cells
(Fig. 4G). These ndings suggest that the low PD-L1 protein level
in H460 cells results from reduced stability of PD-L1 protein.
To investigate whether EpCAM inuences PD-L1 protein stability,
we overexpressed EpCAM in H460 cells (low endogenous PD-L1
expression and no EpCAM expression). We found EpCAM expression
increased PD-L1 protein level, but did not change its mRNA level
(Supplementary Fig. S4A and S4B). A cycloheximide chase assay
conrmed that PD-L1 protein half-life was increased by EpCAM
overexpression (Supplementary Fig. S4C). In addition, knockdown
of EpCAM in H441 cells (high endogenous EpCAM expression)
decreased PD-L1 protein level and protein half-life, but did not change
mRNA level (Fig. 4H and I). Consistent with this idea, H441 cells with
cytomegalovirus promoterdriven overexpression of PD-L1-Myc,
which lack endogenous regulatory elements, such as the PD-L1
promoter, 30untranslated region (UTR), and 50UTR, also exhibited
decreased PD-L1 protein when EpCAM was knocked down (Fig. 4J).
Furthermore, to validate our ndings in samples from human patients
with cancer, PD-L1 and EpCAM protein levels were examined in both
lung and colon tumor specimens. Similar to our results in lung cancer
cell lines, the PD-L1 protein level was highly correlated with EpCAM
expression in a lung tumor tissue array (Fig. 4K). On the other hand,
most of the specimens in the colon tissue array showed only weak
staining for PD-L1, which prevented us from making a clear conclu-
sion, even though the signal for EpCAM was strong (Supplementary
Fig. S5A and S5B). Furthermore, GSEA indicated the related genes of
T-cell activation and proliferation enrichment in low EpCAM expres-
sion of lung cancer specimen (Fig. 4L). Accordingly, we concluded that
EpCAM expression is positively correlated to PD-L1 protein stability.
EpEX stabilizes PD-L1 through the EGFRERK pathway
Regulated intramembrane proteolysis triggers EpCAM-mediated
signal transduction through the dual actions of EpEX shedding by
ADAM17 and EpICD release by presenilin 2containing g-secretase
complex (6). On the basis of the previous studies, EpEX is known to
function as an EGF-like growth factor, triggering EGFR signal-
ing (5, 8, 10). Moreover, activation of the EGFR signaling pathway
can increase PD-L1 expression through various mechan-
isms (17, 19, 21). Thus, we suspected that EpEX might regulate
PD-L1 expression through EGFR signaling. Moreover, EpICD associ-
ates with FHL2 and b-catenin in a complex that translocates to the
nucleus and transcribes downstream genes. Thus, we tested whether
EpEX or EpICD is sufcient to cause PD-L1 stabilization. To test
whether endogenous EpEX or EpICD is necessary for PD-L1 protein
stabilization, ADAM17 inhibitor (TAPI) and g-secretase inhibitor
(DAPT) were utilized to prevent the generation of endogenous EpEX
and EpICD, respectively. Blocking the shedding of endogenous EpEX,
but not EpICD, caused a reduction in PD-L1 protein level. Notably, no
changes in PD-L1 mRNA levels were observed, suggesting that
endogenous EpEX was crucial for PD-L1 protein stability (Fig. 5A).
Moreover, our results show that PD-L1 protein level was increased in a
dose- and time-dependent manner after H441 cells were treated with
EpEX or EGF (Fig. 5B and C). PD-L1 mRNA levels did not increase
after EpEX or EGF treatment (Fig. 5C). These results suggest that
EpEX can increase PD-L1 protein level, but EpICD cannot.
IFNgis known to induce PD-L1 expression through STAT family
mediated transcription (33). We next wondered whether endogenous
EpEX and EpICD also participate in IFNg-mediated PD-L1 upregula-
tion. PD-L1 mRNA was upregulated in H441 cells treated with IFNg
and DMSO, IFNgand TAPI, and IFNgand DAPT (Supplementary
Fig. S6A). Surprisingly, the PD-L1 protein level was not increased as
much in the presence of TAPI as it was in the IFNgand DMSOtreated
cells (Supplementary Fig. S6B). In other words, despite the fact that
PD-L1 mRNA expression was elevated by IFNgstimulation, the
protein level could not be fully upregulated without generation of
EpEX. In addition, PD-L1 protein expression was not impaired in
IFNgand DAPTtreated cells (Supplementary Fig. S6B). While PD-L1
mRNA level was increased in all cells receiving IFNgtreatment, the
protein level was only upregulated in those cells with IFNgand EpEX
or IFNgand EGF treatment (Supplementary Fig. S6C and S6D). Thus,
EpEX appears to be an essential factor for PD-L1 protein upregulation.
The EGFR signaling pathway is important for PD-L1 expression in
cancer cells (16), and our previous study showed that EpEX can induce
EGFR signaling (5). Therefore, we speculated that EpEX might act
through EGFR signaling to prevent PD-L1 degradation. Indeed,
shRNA knockdown of EGFR in H441 cells attenuated EpEX- or
EGF-augmented PD-L1 levels (Fig. 5D). We next investigated which
EGFR downstream signaling pathway was involved in EpEX- and
EGF-mediated upregulation of PD-L1. Treatment with getinib abro-
gated both EpEX- and EGF-induced upregulation of PD-L1 protein
level. Importantly, U0126, which is an MEK inhibitor, but not an AKT
inhibitor, abolished EpEX-mediated upregulation of PD-L1 as well
(Fig. 5E). These results imply that, like EGF, EpEX-mediated upre-
gulation of PD-L1 requires signaling through the MAPK pathway.
Previous studies have shown that both ERK and AKT signaling
could regulate PD-L1 expression through different mechan-
isms (17, 19, 21). Hence, we wanted to know which signaling pathway
was more important for PD-L1 expression in our system. Inhibition of
ERK signaling, but not AKT signaling, decreased PD-L1 expression in
a time-dependent manner (Supplementary Fig. S7A). We then sought
to verify that the MAPK pathway affects PD-L1 protein stability. In the
EpCAM Signaling Regulates Tumorigenesis
AACRJournals.org Cancer Res; 80(22) November 15, 2020 5043
Downloaded from http://aacrjournals.org/cancerres/article-pdf/80/22/5035/2796333/5035.pdf by MD ANDERSON HOSPITAL user on 14 June 2022
Figure 4.
EpCAM is correlated with PD-L1 expression. A, The tumor growth of shLuc or shEpCAM H441 cells in NSG mice with or without PBMC injection. B, Relative tumor size
was evaluated, and weights are shown in the graph. C, Evaluation of CD8
þ
T cells in tumor tissue. Dand E, Western blotting (D) and qRT-PCR analysis (E) of PD-L1 in
H460 and H441 cells. F, Flow cytometry analysis of PD-L1 and EpCAM in H460 and H441 cells. G, H460 and H441 cells were treated with 50 mmol/L cycloheximide
(CHX) for the indicated intervals and analyzed by Western blotting. Protein expression over time is shown in the gra ph. H, The protein and mRNA levels of PD-L1 in
EpCAM-knockdown H441 cells were analyzed by Western blotting and qRT-PCR , respectively. I, EpCAM-knockdown H441 cells were treated with 50 mmol/L
cycloheximide for the indicated intervals, and PD-L1 expression was analyzed by Western blotting. Protein expression over time is shown in the graph. J, The
expression of exogenous PD-L1 in EpCAM-knockdown H441 cells was analyzed by Western blotting. K, IHC staining for EpCAM and PD-L1 was performed in the lung
cancer tissue array, and the positive correlation between EpCAM and PD-L1 expression in patients with lung cancer is shown. L, GSEA enrichment prole of T-cell
activation and proliferation in the lung cancer specimen. Statistical differences were determined by two-tailed Student ttest. N¼6 independent experiments.
Graphical data are shown as mean SEM. ,P<0.01; ,P<0.001.
Chen et al.
Cancer Res; 80(22) November 15, 2020 CANCER RESEARCH
5044
Downloaded from http://aacrjournals.org/cancerres/article-pdf/80/22/5035/2796333/5035.pdf by MD ANDERSON HOSPITAL user on 14 June 2022
Figure 5.
EpEX stabilizes PD-L1 proteinvia EGFRMEK signaling. A, H441 cells were treated with TAPI (ADAM17 inhibitor)or DAPT (g-secretase inhibitor) for 24 hours, and PD-L1
expression was analyzedby Western blotting (left) and qRT-PCR (right). B, H441 cells were treatedwith the indicated concentrationsof EpEX or EGF for 1 hour and the
PD-L1 expression was analyzed by Western blotting.C, After treatment with EpEX (45 nmol/L) or EGF (16.5 nmol/L), for the indicated intervals, PD-L1 expression was
analyzed by Western blotting (left) or qRT-PCR (right).D, EGFR-knockdown H441 cells were treatedwith EpEX or EGF for 1 hour, and PD-L1 expressionwas analyzed
by Western blotting. E, PD-L1 wasdetermined by Western blotting in H441 cells pretreated with getinib, U1026,or wortmannin for 1 hour, followed by treatment with
EpEX or EGF for 1 hour. F, H441 cells were treated with EpAb2-6 or isotype for 16 hours, and PD-L1 expression was analyzed by Western blotting (top). PD-L1
expression was analyzed in EpAb2-6- or isotype-treated H441-derived xenograft tumors (bottom). G, After incubating cells with EpAb2-6 for 16 hours, the protein
half-life of PD-L1 was measured by treatment with 50 mmol/L cycloheximide (CHX) for theindicated intervals andWestern blotting analysis. Protein expression over
time is shown in the graph. H, H441 cells expressingPD-L1-eGFP-Flag were treated with EpAb2-6 for 16 hours and MG132 for 5 hours. The polyubiquitination of PD-L1-
eGFP-Flag was analyzed by Western blotting after immunoprecipitation (IP) with anti-Flag. I, MCF7, BT474, and Cal27 cells were treated with EpAb2-6or isotype for
16 hours, and PD-L1 expression was analyzed by Western blotting. J, After treatment with EpAb2-6 or isotype for 16 hours, H441 cells were cocultured with T cells for
16 hours and the apoptotic cells were quantiedby uorescein Annexin V-FITC/DAPI double staining. IB, immunoblotting. Statistical differences were determined by
two-tailed Student ttest. N¼6 independent experiments. Graphical data are shown as mean SEM. ,P<0.05; ,P<0.01; ,P<0.001.
EpCAM Signaling Regulates Tumorigenesis
AACRJournals.org Cancer Res; 80(22) November 15, 2020 5045
Downloaded from http://aacrjournals.org/cancerres/article-pdf/80/22/5035/2796333/5035.pdf by MD ANDERSON HOSPITAL user on 14 June 2022
presence of U0126 and cycloheximide, the turnover rate of PD-L1 was
higher than it was in the DMSO control (Supplementary Fig. S7B).
Different inhibitors were then used to block the MAPK pathway to
prevent misinterpretation of effects from nonspecic targeting.
Indeed, the different MEK inhibitors, trametinib and U0126, and
ERK inhibitor, SCH772984, all decreased the PD-L1 protein level and
increased polyubiquitination of PD-L1 (Supplementary Fig. S7C and
S7D).
Previous studies have demonstrated that glycosylation of PD-L1 at
N192, N200, and N219 antagonizes GSK3bbinding, which normally
induces phosphorylation-dependent proteasome degradation of PD-
L1 by b-TrCP. As such, nonglycosylated PD-L1 is unstable compared
with glycosylated PD-L1 (17). Consistent with this mechanism, PD-L1
was upregulated in GSK3b-knockdown cells (Supplementary
Fig. S7E). However, EpEX and EGF could both induce PD-L1 upre-
gulation in GSK3b-knockdown cells (Supplementary Fig. S7F), sug-
gesting that GSK3bis dispensable for EpEX- or EGF-mediated PD-L1
stabilization.
We next tested whether EpAb2-6 could attenuate PD-L1 upregula-
tion and EpEX production. After EpAb2-6 treatment, PD-L1 protein
level was downregulated both in vitro and in vivo (Fig. 5F). Consistent
with our other ndings, EpAb2-6 also decreased PD-L1 protein
stability and increased polyubiquitination of PD-L1 (Fig. 5G
and H). Notably, EpAb2-6 treatment inhibited PD-L1 protein level
in lung cancer cells and in cell lines derived from other EpCAM-
positive cancer types, including breast cancer (BT474 and MCF7) and
oral cancer (Cal27; Fig. 5I). Using an apoptosis assay, we further found
that coculturing cancer cells with either EpAb2-6 or T cells could
induce cancer cell apoptosis, and coculturing cancer cells in the
presence of both EpAb2-6 and T cells was more effective in inducing
apoptosis (Fig. 5J).
EpAb2-6 prolongs survival in metastatic and orthotopic mouse
models and improves the efcacy of anti-PD-L1 therapy in the
PBMC CDX model
We next used an animal model of metastatic colon carcinoma to test
whether EpAb2-6 treatment could increase the median overall survival
of metastatic tumor-bearing mice. EpAb2-6 reduced the uorescence
intensity of mouse blood bearing circulating HCT116-GFP cells,
suggesting that EpAb2-6 decreased the number of HCT116-GFP cells
in mouse blood vessels in vivo (Fig. 6A). NOD/SCID mice were
intravenously injected with SW620 cells and then intravenously
treated with EpAb2-6, or an equivalent volume of control IgG,
at 24 and 96 hours after cell injection (antibody was delivered at
20 mg/kg/dose for a total dose of 40 mg/kg). The median survival time
of the EpAb2-6 treatment group was signicantly increased compared
with the control IgG group (Fig. 6B;P<0.005 by the log-rank test).
Because the subcutaneous model may not accurately reproduce
human colon cancer biology (34), an orthotopic mouse model of
colorectal cancer was also established to study the efcacy of our
therapeutic antibody in the colorectal tumor microenvironment. We
investigated the antitumor potential of EpAb2-6 in an orthotopic
model with HCT116-Luc tumors that stably express rey luciferase.
Before the rst therapeutic injection (8 days after tumor cell implan-
tation), growing orthotopic tumors were monitored by biolumines-
cence imaging (Fig. 6C). Mice were treated with either control IgG or
EpAb2-6 (20 mg/kg) every 2 days for 16 days. Notably, the EpAb2-6
treated group showed metastasis at 55 days, while the control group did
so at 45 days (Fig. 6C). Moreover, no signicant changes in body
weight were observed during the treatment period (Supplementary
Fig. S8). At the end of the study, the median survival times for control
IgG and EpAb2-6 treatment groups were 50 and 116.5 days, respec-
tively (Fig. 6D). From these results, we conclude that EpAb2-6
treatment can prolong survival of tumor-bearing mice.
On the basis of our observations that EpAb2-6 treatment decreases
PD-L1 expression both in vitro and in vivo, we further wanted to
evaluate whether EpAb2-6 could improve the therapeutic efcacy of
anti-PD-L1 treatment in vivo. Because EpAb2-6 is not equipped with
cross-reaction with mouse EpCAM, using syngeneic mouse model fails
to evaluate the therapeutic efcacy of EpAb2-6 for immunotherapy. As
illustrated in Fig. 6E, H441 cells were subcutaneously injected into
NSG mice to establish the PBMC-H441xenografted mice model. Two
weeks later, mice were intravenously injected with 10
7
PBMCs and
treated with atezolizumab, an anti-PD-L1 antibody, and EpAb2-6
twice weekly for 1 month. Treatment with atezolizumab or EpAb2-6
alone inhibited tumor growth in PBMC-H441 mice, and combination
therapy produced a much better tumor inhibitory effect (Fig. 6F
and G). At the end of treatment, tumor tissues were collected, and
CD8
þ
T cells were analyzed by ow cytometry. The CD8
þ
T-cell
population was increased in mice receiving combination therapy
(Fig. 6H). Collectively, these data indicate that inhibition of EpCAM
by EpAb2-6 may enhance the efcacy of anti-PD-L1 therapeutics in
PBMC-H441xenografted mice.
Discussion
EpCAM-overexpressing carcinoma cells possess stem celllike
features, and their presence results in high rates of recurrence,
metastasis, and drug resistance (35, 36). We have previously reported
that the EpEX induces EGFR/ERK signaling, which activates ADAM17
and g-secretase to stimulate shedding of EpEX and EpICD. Soluble
EpEX subsequently activates EGFR to form a positive feedback loop,
causing more cleavage of EpCAM. Moreover, the importance of EGFR
signaling on PD-L1 expression has been conclusively shown, and
clinical data demonstrate that PD-L1 expression in the tumor micro-
environment is a determinant of responses to EGFR-targeted thera-
pies. Although EpCAM expression correlates with cancer malignancy,
the molecular mechanism by which EpCAM prevents apoptosis and
suppresses immune response remains unclear.
In this study, we report that soluble EpEX and membrane-bound
EpCAM can directly bind to EGFR through the EGF-like domain I in
EpEX and induce EGFR signaling (Fig. 1). Similarly, recent studies
have reported that EGF-like domains on growth factors could bind to
members of the ErbB/HER family (37) and that membrane-bound
tenascin-C, a matrix component containing EGF-like repeats, could
directly bind to EGFR with very low afnity (38). In addition, because
both EpEX and EpCAM triggered EGFR signaling, the signal may be
transduced in either an autocrine or paracrine manner.
We have previously demonstrated that anti-EpCAM neutralizing
antibody, EpAb2-6, could bind to positions Y95 and D96 of EpCAM to
directly induce cancer cell apoptosis in vitro (23). Inhibition of
EpCAM cleavage or EpEX binding to EGFR through EpAb2-6 might
induce apoptosis both in vitro and in vivo to reduce survival of cancer
stem cells. Notably, we also found that HtrA2 is a direct transcriptional
target of FOXO3a (Fig. 3), based on ChIP evidence indicating that
FOXO3a binds to the HtrA2 promoter upon EpAb2-6 stimulation
(Fig. 3E and F). Interestingly, the expression of XIAP might promote
resistance of colon cancer cells to EpAb2-6 (39), but the functional
outcome of EpAb2-6 treatment was increased apoptosis of cancer cells.
HtrA2 promotes or induces cell death through two different mechan-
isms. One is by directly binding to and inhibiting inhibitor of apoptosis
proteins (IAP), an action that is accompanied by a signicant increase
Chen et al.
Cancer Res; 80(22) November 15, 2020 CANCER RESEARCH
5046
Downloaded from http://aacrjournals.org/cancerres/article-pdf/80/22/5035/2796333/5035.pdf by MD ANDERSON HOSPITAL user on 14 June 2022
in caspase activity. The other mechanism is through a relatively
uncharacterized IAP inhibitionindependent, caspase-independent,
and HtrA2 serine protease activitydependent mechanism (40).
Previous studies have shown that p53 induces the activation of
HtrA2 after oncogenic Ras transformation (41). Oncogenic Ras
increases cytoplasmic p53 accumulation, which promotes p38 MAPK
translocation to mitochondria and phosphorylation of HtrA2 (41).
The phosphorylated HtrA2 releases from mitochondria into the
cytosol and cleaves F-actin to downregulate lamellipodia forma-
tion (41). Proapoptotic protein, Bax, increases endoplasmic reticulum
Ca
þ2
-ATPase inhibitor thapsigargin-induced HtrA2 release from
mitochondria to induce apoptosis in HCT116 cells (42). However,
whether EpAb2-6 regulation of HtrA2 expression involves p53 or Bax
remains poorly understood. To the best of our knowledge, this is the
rst study to demonstrate that inhibition of EpCAM increases HtrA2
gene expression, mitochondria release, and the HtrA2-induced intrin-
sic pathway of apoptosis.
Among all antibody-based immune therapies, anti-PD-1/PD-L1
therapies have the most benecial outcomes in the treatment of various
malignancies. This fact underscores the importance of deeply under-
standing the processes that control PD-L1 expression. We found that
EpCAM-mediated PD-L1 stabilization gives rise to escape from
Figure 6.
EpAb2-6 prolongs median survival in
metastasis and orthotopic colorectal
cancer models and improves the ef-
cacy of anti-PD-L1 therapy in a PBMC
CDX model. A, HCT116 cells transiently
expressing GFP were injected into the
tail veins of SCID mice with EpAb2-6.
Mice in the control group (without
EpAb2-6 treatment) were injected
with the same dose of control IgG.
N¼3 independent experiments. Data
are shown as mean SD. ,P<0.01.
B, NOD/SCID mice were intravenously
injected with 1 10
6
SW620 cells,
followed by treatment with either
control IgG or EpAb2-6 (N¼6).
According to the survival curves, mice
treated with EpAb2-6 exhibited a
greater survival rate than those trea-
ted with control IgG. C, NSG mice were
orthotopically implanted with HCT116-
Luc cells and then treated with control
IgG or EpAb2-6 starting at 8 days after
tumor inoculation. Antibody injections
were performed every 2 days for
16 days, via the tail vein. Tumor growth
was monitored by examining biolumi-
nescence with the IVIS 200 imaging
system. D, KaplanMeier survival plots
and median survival are shown for
both treatment groups (N¼8per
group; P<0.005, log-rank test).
E, A schema for the experiment
to evaluate treatment efcacy of
EpAb2-6 and/or atezolizumab in the
PBMC CDX model. F, The tumor
growth of H441 cells in EpAb2-6-
and/or atezolizumab-treated PBMC
CDX mice. G, Tumor weight was eval-
uated and is shown in the graph.
H, CD8
þ
T cells were quantied in
tumor tissue. Statistical differences
were determined by two-tailed Stu-
dent ttest. N¼6 independent experi-
ments. Graphical data are shown as
mean SEM. ,P<0.05; ,P<0.01;
,P<0.001.
EpCAM Signaling Regulates Tumorigenesis
AACRJournals.org Cancer Res; 80(22) November 15, 2020 5047
Downloaded from http://aacrjournals.org/cancerres/article-pdf/80/22/5035/2796333/5035.pdf by MD ANDERSON HOSPITAL user on 14 June 2022
immune surveillance through the EpEXEGFRERK signaling axis.
According to these ndings, EpCAM might be an excellent option for
combination with anti-PD-1/PD-L1 therapy. Indeed, EpAb2-6
decreases PD-L1 protein level and improves the therapeutic efcacy
of atezolizumab. Thus, our nding reveals a novel action of EpCAM in
the regulation of PD-L1 protein stability and suggests a new strategy of
EpCAM/PD-L1targeted combination therapy. However, we did not
test EpCAM-mediated PD-L1 stabilization or the correlation between
EpCAM and PD-L1 in colon cancer, due to the low PD-L1 expression
in both colon cancer cell lines and tumor specimens. In our model, we
propose that EpCAM can stabilize PD-L1 protein, instead of increas-
ing its transcriptional expression. Therefore, if expression of endog-
enous PD-L1 is already low in a tumor, more EpCAM expression
would not increase its protein level to any major extent. Nevertheless,
more colon cancer cell lines or tumor specimens should be examined to
further test whether EpCAM-mediated PD-L1 stabilization and
EpCAM/PD-L1targeted combination therapy might be benecial in
colon cancer.
Activating EGFR mutations were reported to be associated with
increased PD-L1 expression, and EGFR inhibitors can reduce PD-L1
expression in NSCLC cell lines, suggesting that EGFR signaling can
trigger immune escape (16). Further investigations demonstrated that
EGF induces PD-L1 expression at a posttranslational level, but does
not affect PD-L1 mRNA expression. The authors also found that EGF
Figure 7.
EpCAM signaling regulates tumor progression.
Chen et al.
Cancer Res; 80(22) November 15, 2020 CANCER RESEARCH
5048
Downloaded from http://aacrjournals.org/cancerres/article-pdf/80/22/5035/2796333/5035.pdf by MD ANDERSON HOSPITAL user on 14 June 2022
stabilizes PD-L1 by inducing its glycosylation and inactivating GSK3b,
which promotes degradation of nonglycosylated PD-L1 (17). How-
ever, we found that EpEX-mediated stabilization of glycosylated PD-
L1 occurs through the MAPK pathway. This mechanism differs from
GSK3b-mediated degradation of nonglycosylated PD-L1. Whether the
MAPK pathway regulates stability of nonglycosylated PD-L1 needs
further examination.
While the crystal structure of EpEX has been published previous-
ly (43), the crystal structure of the EpAb2-6EpEX complex is worth
exploring. A crystal structure for EpAb2-6EpEX would reveal the
detailed EpAb2-6 binding sites and conformational changes in EpEX
that induce apoptosis. Previous studies have demonstrated that mAbs
bind to receptors and induce conformational changes in the receptor,
which modulates the expression, recycling, and function of the recep-
tor as well as ligand afnity (44, 45). Previous studies have demon-
strated that structure of EpEX forms a cis-dimer and endocytosis of
EpCAM into acidic compartments dissociates the dimer to permit
cleavage (43, 46, 47). EpAb2-6 might affect the conformation and the
endocytosis of EpCAM to disrupt acidication, dissociation, and/or
cleavage.
To the best of our knowledge, this is the rst study to demonstrate
that inhibition of EpCAM is correlated with an increase in HtrA2 gene
expression. Moreover, this upregulation occurs through FOXO3a and
induces apoptosis. EpEX increases PD-L1 protein stability and the
combination immunotherapy of anti-EpCAM and anti-PD-L1 anti-
bodies provides a novel strategy for cancer therapy (Fig. 7). Most
importantly, our data indicate that the development of therapeutic
antibodies targeting EpCAM may hold great potential in promoting
immune surveillance and eradicating cancer stem cells in the cancer
microenvironment.
Disclosure of Potential Conicts of Interest
No potential conicts of interest were disclosed.
AuthorsContributions
H.-N. Chen: Investigation, methodology, writing-original draft, writing-review
and editing. K.-H. Liang: Investigation, methodology, writing-original draft, writing-
review and editing. J.-K. Lai: Investigation, methodology. C.-H. Lan: Investigation,
methodology. M.-Y. Liao: Investigation, methodology. S.-H. Hung: Investigation,
methodology. Y.-T. Chuang: Investigation, methodology. K.-C. Chen: Investigation,
methodology. W.W.-F. Tsuei: Investigation, methodology. H.-C. Wu:
Conceptualization, data curation, supervision, funding acquisition, validation,
writing-original draft, project administration, writing-review and editing.
Acknowledgments
We thank the Core Facility of the Institute of Cellular and Organismic Biology
(Nankang, Taipei, Taiwan) and National RNAi Core Facility, Academia Sinica
(Nankang, Taipei, Taiwan) for technical support. This research was supported by
Academia Sinica (AS-SUMMIT-108) and the Ministry of Science and Technology
(MOST-108-3114-Y-001-002 and MOST-108-2823-8-001-001 to H.-C. Wu).
The costs of publication of this article were defrayed in part by the payment of page
charges. This article must therefore be hereby marked advertisement in accordance
with 18 U.S.C. Section 1734 solely to indicate this fact.
Received April 18, 2020; revised July 17, 2020; accepted September 22, 2020;
published rst September 25, 2020.
Reference
1. Baeuerle PA, Gires O. EpCAM (CD326) nding its role in cancer. Br J Cancer
2007;96:41723.
2. Patriarca C, Macchi RM, Marschner AK, Mellstedt H. Epithelial cell adhesion
molecule expression (CD326) in cancer: a short review. Cancer Treat Rev 2012;
38:6875.
3. Spizzo G, Went P, Dirnhofer S, Obrist P, Moch H, Baeuerle PA, et al. Over-
expression of epithelial cell adhesion molecule (Ep-CAM) is an independent
prognostic marker for reduced survival of patients with epithelial ovarian cancer.
Gynecol Oncol 2006;103:4838.
4. Spizzo G, Went P, Dirnhofer S, Obrist P, Simon R, Spichtin H, et al. High
Ep-CAM expression is associated with poor prognosis in node-positive breast
cancer. Breast Cancer Res Treat 2004;86:20713.
5. Liang KH, Tso HC, Hung SH, Kuan II, Lai JK, Ke FY, et al. Extracellular domain
of EpCAM enhances tumor progression through EGFR signaling in colon cancer
cells. Cancer Lett 2018;433:16575.
6. Maetzel D, Denzel S, Mack B, Canis M, Went P, Benk M, et al. Nuclear signalling
by tumour-associated antigen EpCAM. Nat Cell Biol 2009;11:16271.
7. Lin CW, Liao MY, Lin WW, Wang YP, Lu TY, Wu HC. Epithelial cell adhesion
molecule regulates tumor initiation and tumorigenesis via activating reprogram-
ming factors and epithelial-mesenchymal transition gene expression in colon
cancer. J Biol Chem 2012;287:3944959.
8. Pan M, Schinke H, Luxenburger E, Kranz G, Shakhtour J, Libl D, et al. EpCAM
ectodomain EpEX is a ligand of EGFR that counteracts EGF- mediated epithelial-
mesenchymal transition through modulation of phospho-ERK1/2 in head and
neck cancers. PLoS Biol 2018;16:e2006624.
9. Lu TY, Lu RM, Liao MY, Yu J, Chung CH, Kao CF, et al. Epithelial cell adhesion
molecule regulation is associated with the maintenance of the undifferentiated
phenotype of human embryonic stem cells. J Biol Chem 2010;285:871932.
10. Liang K-H, Lai J-K, Kuan II, Tso H-C, Wu H-C. EpCAM/EpEX regulate tumor
progression through EGFR signaling in colon cancer cells [abstract]. In:
Proceedings of the American Association for Cancer Research Annual Meeting
2017; 2017 Apr 15; Washington, DC. Philadelphia (PA): AACR; 2017. Abstract
nr 327.
11. Zhang X, Tang N, Hadden TJ, Rishi AK. Akt, FoxO and regulation of apoptosis.
Biochim Biophys Acta 2011;1813:197886.
12. Brunet A, Bonni A, Zigmond MJ, Lin MZ, Juo P, Hu LS, et al. Akt promotes cell
survival by phosphorylating and inhibiting a forkhead transcription factor. Cell
1999;96:85768.
13. TsaiKL,SunYJ,HuangCY,YangJY,HungMC,HsiaoCD.Crystal
structure of the human FOXO3a-DBD/DNA complex suggests the
effects of post-translational modication. Nucleic Acids Res 2007;35:
698494.
14. SkurkC,MaatzH,KimHS,YangJ,AbidMR,AirdWC,etal.TheAkt-
regulated forkhead transcription factor FOXO3a controls endothelial cell
viability through modulation of the caspase-8 inhibitor FLIP. J Biol Chem
2004;279:151325.
15. Fujio Y, Walsh K. Akt mediates cytoprotection of endothelial cells by vascular
endothelial growth factor in an anchorage-dependent manner. J Biol Chem 1999;
274:1634954.
16. Akbay EA, Koyama S, Carretero J, Altabef A, Tchaicha JH, Christensen CL, et al.
Activation of the PD-1 pathway contributes to immune escape in EGFR-driven
lung tumors. Cancer Discov 2013;3:135563.
17. Li CW, Lim SO, Xia W, Lee HH, Chan LC, Kuo CW, et al. Glycosylation and
stabilization of programmed death ligand-1 suppresses T-cell activity.
Nat Commun 2016;7:12632.
18. Sumimoto H, Takano A, Teramoto K, Daigo Y. RAS-mitogen-activated protein
kinase signal is required for enhanced PD-L1 expression in human lung cancers.
PLoS One 2016;11:e0166626.
19. Chen N, Fang W, Zhan J, Hong S, Tang Y, Kang S, et al. Upregulation of PD-L1 by
EGFR activation mediates the immune escape in EGFR-driven NSCLC: impli-
cation for optional immune targeted therapy for NSCLC patients with EGFR
mutation. J Thorac Oncol 2015;10:91023.
20. Fang W, Zhang J, Hong S, Zhan J, Chen N, Qin T, et al. EBV-driven LMP1 and
IFN-gup-regulate PD-L1 in nasopharyngeal carcinoma: implications for onco-
targeted therapy. Oncotarget 2014;5:12189202.
21. Coelho MA, de Carne Trecesson S, Rana S, Zecchin D, Moore C, Molina-Arcas
M, et al. Oncogenic RAS Signaling promotes tumor immunoresistance by
stabilizing PD-L1 mRNA. Immunity 2017;47:108399.
22. Schnell U, Kuipers J, Giepmans BN. EpCAM proteolysis: new fragments with
distinct functions? Biosci Rep 2013;33:e00030.
AACRJournals.org Cancer Res; 80(22) November 15, 2020 5049
EpCAM Signaling Regulates Tumorigenesis
Downloaded from http://aacrjournals.org/cancerres/article-pdf/80/22/5035/2796333/5035.pdf by MD ANDERSON HOSPITAL user on 14 June 2022
23. Liao MY, Lai JK, Kuo MY, Lu RM, Lin CW, Cheng PC, et al. An anti-EpCAM
antibody EpAb2-6 for the treatment of colon cancer. Oncotarget 2015;6:
2494768.
24. Wu CH, Kuo YH, Hong RL, Wu HC. a-Enolase-binding peptide enhances drug
delivery efciency and therapeutic efcacy against colorectal cancer. Sci Transl
Med 2015;7:290ra91.
25. Ning Y, Luo C, Ren K, Quan M, Cao J. FOXO3a-mediated suppression of the self-
renewal capacity of sphere-forming cells derived from the ovarian cancer SKOV3
cell line by 7-diuoromethoxyl-5,40-di-n-octyl genistein. Molecular medicine
reports 2014;9:19828.
26. Lu M, Chen X, Xiao J, Xiang J, Yang L, Chen D. FOXO3a reverses the cisplatin
resistance in ovarian cancer. Arch Med Res 2018;49:848.
27. Yu Y, Guo M, Wei Y, Yu S, Li H, Wang Y, et al. FoxO3a confers cetuximab
resistance in RAS wild-type metastatic colorectal cancer through c-Myc.
Oncotarget 2016;7:80888900.
28. Marzi L, Combes E, Vie N, Ayrolles-Torro A, Tosi D, Desigaud D, et al. FOXO3a
and the MAPK p38 are activated by cetuximab to induce cell death and inhibit
cell proliferation and their expression predicts cetuximab efcacy in colorectal
cancer. Br J Cancer 2016;115:122333.
29. Bhuiyan MS, Fukunaga K. Mitochondrial serine protease HtrA2/Omi as a
potential therapeutic target. Curr Drug Targets 2009;10:37283.
30. Yang Y, Hsu JM, Sun L, Chan LC, Li CW, Hsu JL, et al. Palmitoylation stabilizes
PD-L1 to promote breast tumor growth. Cell Res 2019;29:836.
31. Zhang J, Bu X, Wang H, Zhu Y, Geng Y, Nihira NT, et al. Cyclin D-CDK4 kinase
destabilizes PD-L1 via cullin 3-SPOP to control cancer immune surveillance.
Nature 2018;553:915.
32. Lim SO, Li CW, Xia W, Cha JH, Chan LC, Wu Y, et al. Deubiquitination and
stabilization of PD-L1 by CSN5. Cancer Cell 2016;30:92539.
33. Moon JW, Kong SK, Kim BS, Kim HJ, Lim H, Noh K, et al. IFNginduces PD-L1
overexpression by JAK2/STAT1/IRF-1 signaling in EBV-positive gastric carci-
noma. Sci Rep 2017;7:17810.
34. Taketo MM, Edelmann W. Mouse models of colon cancer. Gastroenterology
2009;136:78098.
35. Liu D, Sun J, Zhu J, Zhou H, Zhang X, Zhang Y. Expression and clinical
signicance of colorectal cancer stem cell marker EpCAM/CD44 in colorectal
cancer. Oncol Lett 2014;7:15448.
36. Singh A, Settleman J. EMT, cancer stem cells and drug resistance: an emerging
axis of evil in the war on cancer. Oncogene 2010;29:474151.
37. Van Zoelen EJ, Stortelers C, Lenferink AE, Van de Poll ML. The EGF domain:
requirements for binding to receptors of the ErbB family. Vitam Horm 2000;59:
99131.
38. Swindle CS, Tran KT, Johnson TD, Banerjee P, Mayes AM, Grifth L, et al.
Epidermal growth factor (EGF)-like repeats of human tenascin-C as ligands for
EGF receptor. J Cell Biol 2001;154:45968.
39. Holcik M, Yeh C, Korneluk RG, Chow T. Translational upregulation of X-linked
inhibitor of apoptosis (XIAP) increases resistance to radiation induced cell death.
Oncogene 2000;19:41747.
40. Suzuki Y, Imai Y, Nakayama H, Takahashi K, Takio K, Takahashi R. A serine
protease, HtrA2, is released from the mitochondria and interacts with XIAP,
inducing cell death. Mol Cell 2001;8:61321.
41. Yamauchi S, Hou YY, Guo AK, Hirata H, Nakajima W, Yip AK, et al. p53-
mediated activation of the mitochondrial protease HtrA2/Omi prevents cell
invasion. J Cell Biol 2014;204:1191207.
42. Yamaguchi H, Bhalla K, Wang HG. Bax plays a pivotal role in thapsigargin-
induced apoptosis of human colon cancer HCT116 cells by controlling Smac/
Diablo and Omi/HtrA2 release from mitochondria. Cancer Res 2003;63:14839.
43. Pavsic M, Guncar G, Djinovic-Carugo K, Lenarcic B. Crystal structure and its
bearing towards an understanding of key biological functions of EpCAM.
Nat Commun 2014;5:4764.
44. Moura IC, Lepelletier Y, Arnulf B, England P, Baude C, Beaumont C, et al. A
neutralizing monoclonal antibody (mAb A24) directed against the transferrin
receptor induces apoptosis of tumor T lymphocytes from ATL patients. Blood
2004;103:183845.
45. Frelinger AL III, Du XP, Plow EF, Ginsberg MH. Monoclonal antibodies to
ligand-occupied conformers of integrin alpha IIb beta 3 (glycoprotein IIb-IIIa)
alter receptor afnity, specicity, and function. J Biol Chem 1991;266:1710611.
46. Hachmeister M, Bobowski KD, Hogl S, Dislich B, Fukumori A, Eggert C, et al.
Regulated intramembrane proteolysis and degradation of murine epithelial cell
adhesion molecule mEpCAM. PLoS One 2013;8:e71836.
47. Tsaktanis T, Kremling H, Pavsic M, von Stackelberg R, Mack B, Fukumori A,
et al. Cleavage and cell adhesion properties of human epithelial cell adhesion
molecule (HEPCAM). J Biol Chem 2015;290:2457491.
Cancer Res; 80(22) November 15, 2020 CANCER RESEARCH
5050
Chen et al.
Downloaded from http://aacrjournals.org/cancerres/article-pdf/80/22/5035/2796333/5035.pdf by MD ANDERSON HOSPITAL user on 14 June 2022
... Through two-step proteolytic processing, EPCAM is sequentially cleaved by TACE and presenilin 2 (PS-2), a protease component of γ-secretase complex, and releases an N-terminal extracellular domain (EpEX) and a 5 kDa C-terminal intracellular domain (EpICD) to initiate signal transduction [33]. To inhibit EPCAM signalling, both TACE inhibitor (TAPI), which block the release of EpEX, and γ-secretase inhibitor (DAPT), which, in turn, block the release of EpICD, are used [23,34]. Significantly more migrated CD8 + T cells were observed in the TAPI + DAPT and TAPI-only groups (Fig. 4a, b). ...
... Finally, EPCAM is reportedly involved in cell proliferation and migration [23,33,34]. Therefore, we investigated the effect of EPCAM expression on cell proliferation and migration in LMS cell lines. ...
... There are few reports on the role of EPCAM in tumour immunity. Moreover, there are only reports of its involvement in PD-L1 protein expression in colorectal cancer and the cytotoxic activity of NK cells in hepatocellular carcinoma [34,45]. This study found that EPCAM can also be used as a target for immunotherapy. ...
Article
Full-text available
Background Leiomyosarcomas are among the most common histological types of soft tissue sarcoma (STS), with no effective treatment available for advanced patients. Lung metastasis, the most common site of distant metastasis, is the primary prognostic factor. We analysed the immune environment targeting lung metastasis of STS to explore new targets for immunotherapy. Methods We analysed the immune environment of primary and lung metastases in 38 patients with STS using immunohistochemistry. Next, we performed gene expression analyses on primary and lung metastatic tissues from six patients with leiomyosarcoma. Using human leiomyosarcoma cell lines, the effects of the identified genes on immune cells were assessed in vitro. Results Immunohistochemistry showed a significant decrease in CD8 ⁺ cells in the lung metastases of leiomyosarcoma. Among the genes upregulated in lung metastases, epithelial cellular adhesion molecule (EPCAM) showed the strongest negative correlation with the number of CD8 ⁺ cells. Transwell assay results showed that the migration of CD8 ⁺ T cells was significantly increased in the conditioned media obtained after inhibition or knock down of EPCAM. Conclusions EPCAM was upregulated in lung metastases of leiomyosarcoma, suggesting inhibition of CD8 ⁺ T cell migration. Our findings suggest that EPCAM could serve as a potential novel therapeutic target for leiomyosarcoma.
... Furthermore, the expression of EpCAM in squamous carcinomas is correlated with increased cellular proliferation and decreased differentiation [8]. Our group previously developed a neutralizing antibody against EpCAM, EpAb2-6, which has strong potential for use as a colorectal carcinoma (CRC) treatment [5,9,10]. Despite its promise as a therapeutic target in CRC, the mechanisms through which EpCAM contributes to tumorigenesis and metastasis are still not completely known. ...
... EpEX contains two epidermal growth factor (EGF)-like domains, and it may serve as a soluble growth factor to activate EGF receptor (EGFR) in the local tumor microenvironment [5,6,9]. A previous report showed that activation of EGFR could trigger RIP of EpCAM to induce EMT [23]. ...
... Orthotopic tumor models were created as previously reported [9]. Briefly, NSCID mice were used for orthotopic implantation of colon cancer cells previously infected with Lenti-Luc virus (lentivirus containing luciferase gene). ...
Article
Full-text available
Background Epithelial cell adhesion molecule (EpCAM) is known to highly expression and promotes cancer progression in many cancer types, including colorectal cancer. While metastasis is one of the main causes of cancer treatment failure, the involvement of EpCAM signaling in metastatic processes is unclear. We propose the potential crosstalk of EpCAM signaling with the HGFR signaling in order to govern metastatic activity in colorectal cancer. Methods Immunoprecipitation (IP), enzyme-linked immunosorbent assay (ELISA), and fluorescence resonance energy transfer (FRET) was conducted to explore the extracellular domain of EpCAM (EpEX) and HGFR interaction. Western blotting was taken to determine the expression of proteins in colorectal cancer (CRC) cell lines. The functions of EpEX in CRC were investigated by proliferation, migration, and invasion analysis. The combined therapy was validated via a tail vein injection method for the metastasis and orthotopic colon cancer models. Results This study demonstrates that the EpEX binds to HGFR and induces downstream signaling in colon cancer cells. Moreover, EpEX and HGF cooperatively mediate HGFR signaling. Furthermore, EpEX enhances the epithelial-to-mesenchymal transition and metastatic potential of colon cancer cells by activating ERK and FAK-AKT signaling pathways, and it further stabilizes active β-catenin and Snail proteins by decreasing GSK3β activity. Finally, we show that the combined treatment of an anti-EpCAM neutralizing antibody (EpAb2-6) and an HGFR inhibitor (crizotinib) significantly inhibits tumor progression and prolongs survival in metastatic and orthotopic animal models of colon cancer. Conclusion Our findings illuminate the molecular mechanisms underlying EpCAM signaling promotion of colon cancer metastasis, further suggesting that the combination of EpAb2-6 and crizotinib may be an effective strategy for treating cancer patients with high EpCAM expression.
... The combination of EpAb2-6 with Atezolizumab, an anti-PD-L1 antibody, has led to near-complete tumor eradication in preclinical models, thereby revealing a synergistic potential that warrants further exploration. Such combination strategies, which capitalize on EpCAM expression in CRC tumors, could represent a significant advance in cancer immunotherapy [45,46]. Collectively, these findings underscore the promise of EpCAM-associated CAR T-cell therapy and associated immunotherapeutic strategies in the treatment of CRC, emphasizing the importance of personalized and targeted approaches in the future of cancer therapy. ...
Article
Full-text available
The epithelial cell adhesion molecule (EpCAM) is a critical glycoprotein involved in cell cycle progression, proliferation, differentiation, migration, and immune evasion. Its role as a target for bispecific antibodies has shown promise in annihilating cancer cells. EpCAM's potential as a biomarker for tumor-initiating cells, characterized by self-renewal and tumorigenic capabilities, underscores its value in early cancer detection, immunotherapy, and targeted drug delivery. While EpCAM monotherapies have been met with limited success, bispecific antibodies targeting both EpCAM and other proteins have exhibited encouraging results in colorectal cancer (CRC) research. The integration of EpCAM-directed nanotechnology in drug delivery systems has emerged as a pivotal innovation in CRC treatment. Moreover, developing chimeric antigen receptor (CAR) T-cell and CAR natural killer (NK) cell therapies opens promising therapeutic avenues for EpCAM-positive CRC patients. Although preliminary, this review sets the stage for future advances. Additionally, this study advances our understanding of the role of non-coding RNAs in CRC, which may be pivotal in gene regulation and could provide insights into the molecular underpinning. The findings suggest that lncRNA, miRNA, and circRNA could serve as novel therapeutic targets or biomarkers, further enriching the landscape of CRC diagnostics and therapeutics.
... Interestingly, a RT-qPCR analysis of the tumor tissue of surviving mice showed a significant decrease in EpCAM expression after Bengamide II treatment (Fig. 6E). In NSCLC, elevated expression of EpCAM has been associated with unfavorable overall survival outcomes [41] and tumor progression [42]. EpCAM expression is also correlated with CD44 and CD166, and the presence of the triple-positive phenotype (EpCAM+/CD44 +/CD166 +) in NSCLC indicates higher self-renewal capacity, clonal heterogeneity, and gene expression associated with stemness [43]. ...
Article
Full-text available
Lung cancer is the most commonly diagnosed cancer and the one that causes the most deaths worldwide, so there is a need for therapies that improve survival rates. Products derived from marine organisms are a source of novel and potent antitumor compounds, but they present the great obstacle of their obtaining from the natural environment and the problems associated with the synthesis and biological effects of chemical analogues. In this work, a Bengamide analogue (Bengamide II) was chemically synthesized and in vitro and in vivo studies were performed to determine its antitumor activity and mechanisms of action. It was shown to have potent anti-proliferative activity in lung cancer lines in 2D and 3D models. In addition, Bengamide II-treated cells showed G2/M and G0/G1 cell cycle arrest, together with a decrease in the proliferation marker Ki67. As for the mechanism of action, the treatment was associated with increased LC3-II expression and production of acidic vesicles signaling autophagy. In addition, Bengamide II treatment was associated with caspase-3 activation and DNA fragmentation related to apoptosis. Furthermore, a reduction of VEGFA expression, related to angiogenesis, was also observed. In vivo studies showed that Bengamide II markedly reduced tumor volume and metastases increasing survival. Additionally, it revealed no systemic toxicity in in vivo models at the therapeutic doses used, which is essential for its future clinical use. Taken together, the chemically synthesized bengamide analogue Bengamide II, is a promising drug for lung cancer treatment showing relevant antitumor activity and significant safety.
Article
Full-text available
Cancer stem cells (CSCs) represent a subpopulation within tumors that promote cancer progression, metastasis, and recurrence due to their self-renewal capacity and resistance to conventional therapies. CSC-specific markers and signaling pathways highly active in CSCs have emerged as a promising strategy for improving patient outcomes. This review provides a comprehensive overview of the therapeutic targets associated with CSCs of solid tumors across various cancer types, including key molecular markers aldehyde dehydrogenases, CD44, epithelial cellular adhesion molecule, and CD133 and signaling pathways such as Wnt/β-catenin, Notch, and Sonic Hedgehog. We discuss a wide array of therapeutic modalities ranging from targeted antibodies, small molecule inhibitors, and near-infrared photoimmunotherapy to advanced genetic approaches like RNA interference, CRISPR/Cas9 technology, aptamers, antisense oligonucleotides, chimeric antigen receptor (CAR) T cells, CAR natural killer cells, bispecific T cell engagers, immunotoxins, drug-antibody conjugates, therapeutic peptides, and dendritic cell vaccines. This review spans developments from preclinical investigations to ongoing clinical trials, highlighting the innovative targeting strategies that have been informed by CSC-associated pathways and molecules to overcome therapeutic resistance. We aim to provide insights into the potential of these therapies to revolutionize cancer treatment, underscoring the critical need for a multi-faceted approach in the battle against cancer. This comprehensive analysis demonstrates how advances made in the CSC field have informed significant developments in novel targeted therapeutic approaches, with the ultimate goal of achieving more effective and durable responses in cancer patients.
Article
Background Epithelial cell adhesion molecule (EpCAM), a well-established marker for circulating tumor cells, plays a crucial role in the complex process of cancer metastasis. The primary objective of this investigation is to study EpCAM expression in pan-cancer and elucidate its significance in the context of kidney renal clear cell carcinoma (KIRC). Methods Data obtained from the public database was harnessed for the comprehensive assessment of the EpCAM expression levels and prognostic and clinicopathological correlations in thirty-three types of cancer. EpCAM was validated in our own KIRC sequencing and immunohistochemical cohorts. Subsequently, an in-depth exploration was conducted to scrutinize the interrelationship between EpCAM and various facets, including immune cells, immune checkpoints, and chemotherapy drugs. We employed Cox regression analysis to identify prognostic immunomodulators associated with EpCAM, which were subsequently utilized in the development of a prognostic model. The model was validated in our own clinical cohort and public datasets, and compared with 137 published models. The role of EpCAM in KIRC was explored by biological function experiments in vitro. Results While EpCAM exhibited pronounced overexpression across a wide spectrum of cancer types, a notable reduction was observed in KIRC tissues. As grade increased, EpCAM expression decreased. EpCAM expression decreased in patients without metastasis. EpCAM mRNA and protein levels were used as independent, favorable prognostic factors in patients with KIRC in our own cohort. The expression of EpCAM exhibited strong associations with immune-related pathways, demonstrating an inverse correlation with the majority of immune cell types. Immune checkpoint inhibitors exert better therapeutic effects on patients with low EpCAM expression. In addition, EpCAM can be used as a drug resistance indicator and guide the clinical medication of patients with KIRC. A robust model, which had good predictive accuracy and applicability, showed significant superiority over other models. Importantly, EpCAM played the dual roles of promoting proliferation and resisting metastasis in KIRC. Conclusion In the context of KIRC, EpCAM assumes a surprising dual role, where it not only facilitates cell proliferation but also exerts resistance against the metastatic process. EpCAM serves as a standalone prognostic marker for patients with KIRC, and related models can also effectively predict prognosis. These discoveries offer novel perspectives on the functional significance of EpCAM in the context of KIRC.
Article
Full-text available
Simple Summary Aptamers are versatile molecules for cancer treatment. They can be used to target specific molecules, conjugated to additional nucleic acid moieties or payloads. In this review, we will summarize the advancements and challenges related to the use of aptamers for the treatment of ovarian cancer. We will specifically focus on aptamers as therapeutics alone or conjugated to non-coding RNA, nanoparticles, and multivalent aptamers, highlighting their potential as therapeutics by providing an overview of the latest advancements in the field reported in the literature and discussing the main challenges that need to be overcome to improve the effectiveness of aptamers for the treatment of ovarian cancer. Abstract Ovarian cancer (OC) is the most common lethal gynecologic cause of death in women worldwide, with a high mortality rate and increasing incidence. Despite advancements in the treatment, most OC patients still die from their disease due to late-stage diagnosis, the lack of effective diagnostic methods, and relapses. Aptamers, synthetic, short single-stranded oligonucleotides, have emerged as promising anticancer therapeutics. Their ability to selectively bind to target molecules, including cancer-related proteins and receptors, has revolutionized drug discovery and biomarker identification. Aptamers offer unique insights into the molecular pathways involved in cancer development and progression. Moreover, they show immense potential as drug delivery systems, enabling targeted delivery of therapeutic agents to cancer cells while minimizing off-target effects and reducing systemic toxicity. In the context of OC, the integration of aptamers with non-coding RNAs (ncRNAs) presents an opportunity for precise and efficient gene targeting. Additionally, the conjugation of aptamers with nanoparticles allows for accurate and targeted delivery of ncRNAs to specific cells, tissues, or organs. In this review, we will summarize the potential use and challenges associated with the use of aptamers alone or aptamer–ncRNA conjugates, nanoparticles, and multivalent aptamer-based therapeutics for the treatment of OC.
Article
Full-text available
Head and neck squamous cell carcinomas (HNSCCs) are characterized by outstanding molecular heterogeneity that results in severe therapy resistance and poor clinical outcome. Inter- and intratumoral heterogeneity in epithelial-mesenchymal transition (EMT) was recently revealed as a major parameter of poor clinical outcome. Here, we addressed the expression and function of the therapeutic target epidermal growth factor receptor (EGFR) and of the major determinant of epithelial differentiation epithelial cell adhesion molecule (EpCAM) in clinical samples and in vitro models of HNSCCs. We describe improved survival of EGFRlow/EpCAMhigh HNSCC patients (n = 180) and provide a molecular basis for the observed disparities in clinical outcome. EGF/EGFR have concentration-dependent dual capacities as inducers of proliferation and EMT through differential activation of the central molecular switch phosphorylated extracellular signal–regulated kinase 1/2 (pERK1/2) and EMT transcription factors (EMT-TFs) Snail, zinc finger E-box-binding homeobox 1 (Zeb1), and Slug. Furthermore, soluble ectodomain extracellular domain of EpCAM (EpEX) was identified as a ligand of EGFR that activates pERK1/2 and phosphorylated AKT (pAKT) and induces EGFR-dependent proliferation but represses EGF-mediated EMT, Snail, Zeb1, and Slug activation and cell migration. EMT repression by EpEX is realized through competitive modulation of pERK1/2 activation strength and inhibition of EMT-TFs, which is reflected in levels of pERK1/2 and its target Slug in clinical samples. Accordingly, high expression of pERK1/2 and/or Slug predicted poor outcome of HNSCCs. Hence, EpEX is a ligand of EGFR that induces proliferation but counteracts EMT mediated by the EGF/EGFR/pERK1/2 axis. Therefore, the emerging EGFR/EpCAM molecular cross talk represents a promising target to improve patient-tailored adjuvant treatment of HNSCCs.
Article
Full-text available
Programmed death-ligand 1 (PD-L1) acts as an immune checkpoint inhibitor in various cancers. PD-L1 is known to be more frequently expressed in EBV (+) gastric cancer (GC). However, the mechanisms underlying the regulation of PD-L1 expression in EBV (+) GC remain unclear. We investigated the basal and inducible PD-L1 expressions in GC cells. PD-L1 expression was upregulated upon treatment with IFNγ in both EBV (-) and EBV (+) GC cells. Upon stimulation with the same concentration of IFNγ for 24 h, EBV (+) SNU-719 cells showed dramatically higher PD-L1 expression levels by activating JAK2/STAT1/IRF-1 signaling than those of EBV (-) AGS cells. PD-L1 promoter assays, chromatin immunoprecipitation, and electrophoretic mobility shift assays revealed that IFNγ-inducible PD-L1 overexpression is primarily mediated by the putative IRF-1α site of the PD-L1 promoter in EBV (+) SNU-719 cells. Moreover, EBNA1 knockdown reduced both constitutive and IFNγ-inducible PD-L1 promoter activity by decreasing the transcript and protein levels of JAK2 and subsequently STAT1/IRF-1/PD-L1 signaling. EBNA1 is suggested to be moderately enhance both constitutive and IFNγ-inducible PD-L1 expression in EBV (+) GC cells. Thus, the signaling proteins and EBNA1 that regulate PD-L1 expression are potential therapeutic targets in EBV (+) GC.
Article
Full-text available
The immunosuppressive protein PD-L1 is upregulated in many cancers and contributes to evasion of the host immune system. The relative importance of the tumor microenvironment and cancer cell-intrinsic signaling in the regulation of PD-L1 expression remains unclear. We report that oncogenic RAS signaling can upregulate tumor cell PD-L1 expression through a mechanism involving increases in PD-L1 mRNA stability via modulation of the AU-rich element-binding protein tristetraprolin (TTP). TTP negatively regulates PD-L1 expression through AU-rich elements in the 3' UTR of PD-L1 mRNA. MEK signaling downstream of RAS leads to phosphorylation and inhibition of TTP by the kinase MK2. In human lung and colorectal tumors, RAS pathway activation is associated with elevated PD-L1 expression. In vivo, restoration of TTP expression enhances anti-tumor immunity dependent on degradation of PD-L1 mRNA. We demonstrate that RAS can drive cell-intrinsic PD-L1 expression, thus presenting therapeutic opportunities to reverse the innately immunoresistant phenotype of RAS mutant cancers.
Article
Full-text available
Treatments that target immune checkpoints, such as the one mediated by programmed cell death protein 1 (PD-1) and its ligand PD-L1, have been approved for treating human cancers with durable clinical benefit1,2. However, many cancer patients fail to respond to anti-PD-1/PD-L1 treatment, and the underlying mechanism(s) is not well understood3–5. Recent studies revealed that response to PD-1/PD-L1 blockade might correlate with PD-L1 expression levels in tumor cells6,7. Hence, it is important to mechanistically understand the pathways controlling PD-L1 protein expression and stability, which can offer a molecular basis to improve the clinical response rate and efficacy of PD-1/PD-L1 blockade in cancer patients. Here, we report that PD-L1 protein abundance is regulated by cyclin D-CDK4 and the Cullin 3SPOP E3 ligase via proteasome-mediated degradation. Inhibition of CDK4/6 in vivo elevates PD-L1 protein levels, largely by inhibiting cyclin D–CDK4-mediated phosphorylation of SPOP and thereby promoting SPOP degradation by APC/CCdh1. Loss-of-function mutations in SPOP compromise ubiquitination-mediated PD-L1 degradation, leading to increased PD-L1 levels and reduced numbers of tumor-infiltrating lymphocytes (TILs) in mouse tumors and in primary human prostate cancer specimens. Notably, combining CDK4/6 inhibitor treatment with anti-PD-1 immunotherapy enhances tumor regression and dramatically improves overall survival rates in mouse tumor models. Our study uncovers a novel molecular mechanism for regulating PD-L1 protein stability by a cell cycle kinase and reveals the potential for using combination treatment with CDK4/6 inhibitors and PD-1/PD-L1 immune checkpoint blockade to enhance therapeutic efficacy for human cancers.
Article
Epithelial cell adhesion molecule (EpCAM) is highly expressed in colon cancers, but its role in cancer progression remains to be elucidated. In this work, we found that the extracellular domain of EpCAM (EpEX) activated EGFR and downstream ERK1/2 signaling to promote colon cancer cell migration and proliferation, as well as tumor growth. Mechanistically, we discovered that EpEX-EGFR-ERK1/2 signaling positively regulated intramembrane proteolysis (RIP) of EpCAM and shedding of the intracellular domain (EpICD). Treatment with an EGFR inhibitor ablated the EpEX-induced phosphorylation of ERK1/2 and AKT. Additionally, treatment with inhibitors of either EGFR or MEK decreased EpEX-induced EpICD shedding and further revealed that EpICD is necessary for nuclear accumulation of β-catenin and the induction of HIF1α target gene expression in vitro and in vivo. Moreover, an anti-EpCAM neutralizing monoclonal antibody, EpAb2-6, inhibited the nuclear translocation of EpICD and β-catenin and induced apoptosis in colon cancer cells. Importantly, analysis of colorectal cancer tissues showed that nuclear accumulation of EpICD was highly correlated with metastasis and poor prognosis, suggesting that it may play an important functional role in cancer progression. Thus, we provide novel insights into the mechanisms and functions of EpEX-mediated signaling, which may be considered as a promising target for the treatment of colon cancer.
Article
Objective: Ovarian cancer is one of the most serious disease in female reproductive system. Platinum is the first-line drug for the treatment of ovarian cancer, while the resistance of platinum drug in clinical hindered the relief ovarian cancer. Our previous study found that decreased FOXO3a might be a poor prognosis in human ovarian cancer. In this research, we study whether FOXO3a was involved in the mechanism of platinum drug resistance. Methods: The CCK-8 and FACS analysis were used to monitor the survival of ovarian cancer, and the FOXO3a expression was detected by western-blot. Results: We found that FOXO3a expression upregulated significantly in A2780 compared with A2780/DDP cells with the treatment of platinum. Moreover, overexpression of FOXO3a in ovarian cancer inversed the platinum resistance in ovarian cancer. Conclusion: These observations reminded that the role of FOXO3a might be one of the critical mechanisms in developing platinum drug resistance in ovarian cancer.
Article
Epithelial cell adhesion molecule (EpCAM) is highly expressed in advanced epithelial cancers and tumor-initiated cells (TICs), but its roles in cancer progression remain to be elucidated. Here, we showed that the extracellular domain of EpCAM (EpEX) could bind to EGFR through EGF-like domain I, and subsequently activated its downstream molecules, ERK1/2 and Akt. EGFR inhibitor and knockdown of EGFR by shRNA ablated EpEX-induced ERK1/2 phosphorylation. Regulated intramembrane proteolysis (RIP) of EpCAM was induced similarly by EpEX and EGF through EGFR-dependent activation of ERK pathway. MEK inhibitor, U0126, could abolish ADAM17 and PS2 phosphorylation induced by EpEX. EpAb2-6, an anti-EpEX neutralizing monoclonal antibody, inhibits EpEX-activated EGFR-PI3K-AKT pathway in detached colon cancer cells. Moreover, intracellular domain of EpCAM (EpICD), the product of RIP-induced cleavage of EpCAM, is necessary for nuclear accumulation of β-catenin, and their target gene expressions in vitro and in mouse xenograft models. We also found that an increase of nuclear EpICD observed in CRCs predicted metastasis and poor prognosis in CRC patients. Finally, in animal model studies, EpAb2-6 therapy exhibited an enhanced antitumor effect and markedly extended the survival time of mice with human colorectal cancer in metastatic and orthotopic models. These results demonstrate that EpEX works as a growth factor in activating EGFR-mediated signaling, and as a potential target for treatment of colon cancer. This research was supported by grants from Academia Sinica and Ministry of Science and Technology [MOST 104-0210-01-09-02,MOST 105-0210-01-13-01], and the National Science Council (NSC103-2321-B-001-064), Taiwan (to H-C Wu). Citation Format: Kang-Hao Liang, Jun-Kai Lai, I-I Kuan, Hsien-Cheng Tso, Han-Chung Wu. EpCAM/EpEX regulate tumor progression through EGFR signaling in colon cancer cells [abstract]. In: Proceedings of the American Association for Cancer Research Annual Meeting 2017; 2017 Apr 1-5; Washington, DC. Philadelphia (PA): AACR; Cancer Res 2017;77(13 Suppl):Abstract nr 327. doi:10.1158/1538-7445.AM2017-327
Article
Pro-inflammatory cytokines produced in the tumor microenvironment lead to eradication of anti-tumor immunity and enhanced tumor cell survival. In the current study, we identified tumor necrosis factor alpha (TNF-α) as a major factor triggering cancer cell immunosuppression against T cell surveillance via stabilization of programmed cell death-ligand 1 (PD-L1). We demonstrated that COP9 signalosome 5 (CSN5), induced by NF-κB p65, is required for TNF-α-mediated PD-L1 stabilization in cancer cells. CSN5 inhibits the ubiquitination and degradation of PD-L1. Inhibition of CSN5 by curcumin diminished cancer cell PD-L1 expression and sensitized cancer cells to anti-CTLA4 therapy.