ArticlePDF Available

MiR-126 inhibits proliferation of small cell lung cancer cells by targeting SLC7A5

Wiley
FEBS Letters
Authors:
  • Biological Research Centre

Abstract and Figures

Despite intensive efforts to improve therapies, small cell lung cancer (SCLC) still has a dismal median survival of 18 months. Since miR-126 is under-expressed in the majority of SCLC tumors, we investigated the effect of miR-126 overexpression on the proliferation and cell cycle distribution of H69 cells. Our results demonstrate that miR-126 inhibits proliferation of H69 cells, by delaying the cells in the G1 phase. Short interfering RNA (siRNA) mediated suppression of SLC7A5, a predicted target of mir-126, has the same effect on H69 cells. We also show for the first time that SLC7A5 is a direct target of miR-126.
Content may be subject to copyright.
miR-126 inhibits proliferation of small cell lung cancer cells by targeting SLC7A5
Edit Miko
a
, Zoltán Margitai
b
, Zsolt Czimmerer
a
, Ildikó Várkonyi
c
, Balázs Dezs}
o
c
, Árpád Lányi
d
,
Zsolt Bacsó
b
, Beáta Scholtz
a,
a
Dept. of Biochemistry and Molecular Biology, Clinical Genomics Center, University of Debrecen Medical and Health Science Center (UDMHSC), Hungary
b
Dept. of Biophysics and Cell Biology, UDMHSC, Hungary
c
Dept. of Pathology, UDMHSC, Hungary
d
Institute of Immunology, UDMHSC, Hungary
article info
Article history:
Received 30 December 2010
Revised 16 March 2011
Accepted 16 March 2011
Available online 23 March 2011
Edited by Tamas Dalmay
Keywords:
Small cell lung cancer
miR-126
SLC7A5
G1 delay
abstract
Despite intensive efforts to improve therapies, small cell lung cancer (SCLC) still has a dismal median
survival of 18 months. Since miR-126 is under-expressed in the majority of SCLC tumors, we inves-
tigated the effect of miR-126 overexpression on the proliferation and cell cycle distribution of H69
cells. Our results demonstrate that miR-126 inhibits proliferation of H69 cells, by delaying the cells
in the G1 phase. Short interfering RNA (siRNA) mediated suppression of SLC7A5, a predicted target of
mir-126, has the same effect on H69 cells. We also show for the first time that SLC7A5 is a direct tar-
get of miR-126.
Ó
2011 Federation of European Biochemical Societies. Published by Elsevier B.V. All rights reserved.
1. Introduction
Small cell lung cancer (SCLC) is an aggressive form of lung can-
cer, and it appears that current therapeutical regimens have
reached their maximum potential for this type of cancer. Therefore,
development of targeted therapies for SCLC seems to be inevitable
to improve treatment. SCLC is a fast-growing tumor, whose prolif-
eration is being fuelled in part by several aberrantly activated pro-
proliferative signalling pathways. The role of microRNAs (miRNAs)
in regulating cell proliferation in normal and tumor cells is well
established. In addition, due to the nature of miRNA–mRNA inter-
actions, the same miRNA may regulate a given pro-proliferative
pathway through multiple proteins, or may affect parallel path-
ways through different targets. miR-126 appears to have a complex
role in regulating cellular proliferation. It has an anti-proliferative
effect in several tumor types including non-small cell lung cancer
(NSCLC) cells and colon cancer cells, through targeting different
members of the PI3K/Akt pathway [1–3], or in breast cancer cells,
by targeting IRS1 [4]. In addition to its functions regulating the cell
cycle, miR-126 is a key regulator of vessel development, by target-
ing SPRED1 and PIK3R2 in endothelial cells [5,6]. Lastly, miR-126
may be involved in the metastatic process, as evidenced by its
effects on mammary or gastric carcinoma cell migration [7–9].
Since our previous work has shown that miR-126 is uniformly
under-expressed in primary SCLC tumors [10], in the present study
we investigated its role in regulating the proliferation of SCLC cells.
2. Materials and methods
2.1. Cell culture and transfection
Human H69 and HTB-172 SCLC cell lines were purchased from
the American Type Culture Collection. The cells were grown in
RPMI 1640 supplemented with 10% fetal bovine serum (FBS) and
penicillin/streptomycin at 37 °C with 5% CO
2
. miRNA Precursor
Molecule hsa-miR-126 (PM12841) and Pre-miR Negative
Control 1 (AM17110) were obtained from Ambion
(Austin,Texas). Short interfering RNAs (siRNAs) targeting SLC7A5
(s15653), PLK2 (s64) and Silencer Select Negative Control 1 siR-
NA (4390843) were obtained from Applied Biosystems (Foster Ci-
ty,CA). Cells were transfected with pre-miR-126, SLC7A5, PLK2
siRNAs and controls at a final concentration of 50 nM in all exper-
iments, using Lipofectamine 2000 (Invitrogen). Cells were incu-
bated with the transfection complexes for 6 h before replacing
the medium. Cells were refed daily with fresh growth medium.
0014-5793/$36.00 Ó2011 Federation of European Biochemical Societies. Published by Elsevier B.V. All rights reserved.
doi:10.1016/j.febslet.2011.03.039
Abbreviations: SCLC, small cell lung cancer; ECM, extracellular matrix; miRNA,
microRNA; NSCLC, non-small cell lung cancer; AML, acute myeloid leukemia;
siRNA, short interfering RNA; qPCR, real-time quantitative polymerase chain
reaction; UTR, untranslated region; RNAi, RNA interference
Corresponding author. Address: Dept. of Biochemistry and Molecular Biology,
H-4010 Debrecen, Egyetem tér 1, Hungary. Fax: +36 52 314 989.
E-mail address: scholtz@dote.hu (B. Scholtz).
FEBS Letters 585 (2011) 1191–1196
journal homepage: www.FEBSLetters.org
2.2. Establishing the growth curves
H69 cells: 1.2 10
6
cell/well were plated in 6-well plates and
transfected with miRNA precursors. Forty-eight hours post-
transfection 1/10th of the cells were replated in 24-well plates,
and grown further. Cell numbers were determined by trypan blue
exclusion at different time points (48, 72, 96 h) post-transfection.
Results represent the mean of three independent experiments.
Transfection and proliferation analysis of HTB-172 cells is
described in Supplementary figure legends,Fig. 1.
2.3. Cell cycle analysis
1.2 10
6
cell/well were plated in 6-well plates and transfected
with pre-miRs and siRNAs. Forty-eight hours post-transfection 1/
5th of the cells from each transfection were replated and grown fur-
ther. Cells were harvested at 72 and 96 h post-transfection, washed
in PBS and fixed in ice-cold 95% methanol at 20 °C. Fixed cells were
washed twice in PBS, then resuspended in 0.5 ml PBS containing
propidium iodide (50
l
g/ml) and RNase A (200
l
g/ml) and were
stained overnight at 4 °C. Measurements were performed on a FACS-
Array 96-well plate flow cytometric bioanalyzer (Becton Dickinson).
The DNA dye was excited with the 532 nm laser line and emission
was collected in the yellow channel in linear mode with a 585/
42 nm bandpass filter. Cell clusters were gated out using FSC-A/
SSC-A (Area) and FSC-W/SSC-W (Width) 2D-histograms. Fluores-
cence intensity data were fit with the automatic one cycle diploid
model of the Modfit LT 3.0 software (Verity Software House) with
the AutoDebris compensation, AutoAggregate Compensation and
Apoptosis Model turned on. In the measurements the G1–G2 linear-
ity ratio was around 1.92 and the R.C.S. of the fit (reduced
v
2
,a
measure of goodness of fit) was less than 5. All samples were
prepared in triplicates and generally 50 000 cells were collected
from each well.
2.4. Total RNA extraction and qRT-PCR
Total RNA was isolated using Trizol reagent (Invitrogen), fol-
lowed by DNase I treatment (Ambion). For quantification of
mRNAs, mature miR-126 and RNU38B, reverse transcription was
performed by using High-Capacity cDNA Reverse Transcription
Kit (Applied Biosystems). RT primers for mature miR-126 and
RNU38B were supplied by Applied Biosystems. Real-time quantita-
tive polymerase chain reaction (qPCR) was performed with Fast-
Start SYBR Green Master Mix (Roche) with 0.3
l
M of forward
and reverse primers on an ABI 7900 HT sequence detection system.
The PCR program used for amplification was: 95 °C for 10 min, fol-
lowed by 40 cycles of 95 °C for 15 s, 60 °C for 30 s, 72 °C for 45 s.
HPRT1 was used for normalization. PCR primer sequences are in
Supplementary Table 1. For semi-quantitative PCR of mature
miR-126 and RNU38B, specific primers were supplied by Applied
Biosystems. The PCR program used for semi-quantitative amplifi-
cation was: 95 °C for 10 min, followed by 15 cycles of 15 s at
95 °C and 60 s at 60 °C. PCR products were separated by agarose
gel electrophoresis and visualized by EtBr staining.
2.5. Western blot
Cells were washed with PBS and lysed in 2X SDS loading buffer.
Proteins were separated on 10% SDS polyacrylamide gels and
transferred onto nitrocellulose membrane by electroblotting.
Membranes were blocked with 2.5% nonfat milk, and probed
with primary antibody against human SLC7A5 (3157-1, 1:1000,
Epitomics), PLK2 (Snk/H90, sc-25421, 1:100, Santa Cruz
Fig. 1. Effects of miR-126 overexpression on the proliferation of H69 cells. H69 cells were transfected with pre-miR-126 or NegmiR (negative control). Asterisks indicate
significant t-test results (P< 0.05). (A) MiR-126 overexpression inhibits proliferation of H69 cells. Following transfection, viable cells were counted by trypan blue exclusionat
the indicated time points. (B) Representative cell cycle analysis of H69 cells at 96 h post-transfection, analyzed on a FACSArray bioanalyzer. (C) MiR-126 overexpression
delays H69 cells in G1 phase of the cell cycle. Cells were analyzed on a FACSArray bioanalyzer at 72 or 96 h post-transfection.
1192 E. Miko et al. / FEBS Letters 585 (2011) 1191–1196
Biotechnology) and GAPDH (6C5, sc-32233, 1:1000, Santa Cruz
Biotechnology). The membranes were further probed with
horseradish peroxidase-conjugated secondary antibodies
(1:10,000; anti-mouse or anti-rabbit; Amersham) and proteins
were visualized by SuperSignal West Pico chemiluminescent
substrate (Pierce).
2.6. SLC7A5 3
0
untranslated region (UTR) cloning and luciferase
reporter assay
For luciferase reporter assays, 331 bp from 3
0
UTR of SLC7A5
gene, including the miR-126 target site, was amplified by PCR
using F1 and R1 primers with XhoI and NotI sites. PCR was per-
formed on H69 cDNA created by High-Capacity cDNA Reverse
Transcription Kit (Applied Biosystems). The XhoI/NotI-digested
PCR product was cloned into the XhoI/NotI-digested psiCHECK2
dual luciferase vector (Promega). F1/R2 and F2/R1 primers were
used to delete the miR-126 target site from the 3
0
UTR. After mixing
the two PCR products, and digestion with XhoI and NotI, the 3
0
UTR
fragment with a deleted miR-126 binding site was cloned into
XhoI/NotI-digested psiCHECK2 vector. Primer sequences are given
in Supplementary Table 2. H69 cells were cotransfected with
500 ng psiCHECK2 constructs (WT-UTR or DEL-UTR) and 50 nM
pre-miR-126 or pre-miR-negative control in 6-well plates. Forty-
eight hours post-transfection, firefly and renilla luciferase activities
were measured using Dual-Glo Luciferase assay system (Promega)
in a Perkin Elmer Victor3V Multilabel Plate Reader. The ratio of the
luminescent signals from renilla versus firefly was used to deter-
mine the target specificity of miR-126. All experiments were done
in triplicate.
2.7. Statistical analysis
Statistical analysis was done using GraphPad Prism IV software.
Pvalues were calculated by paired t-test. Pvalues <0.05 were
considered significant.
2.8. Immunohistochemistry
Tissue microarrays of formalin-fixed paraffin-embedded surgical
specimens representing primary SCLC tumors were constructed.
These tumor specimens were characterized in our previous study
[8] as chromogranin-A/synaptophysin and thyroid-transcription fac-
tor-1 (TTF1) positive tumors (>70%). Following hematoxylin-eosin
staining, serial sections and antigen retrieving were made for immu-
nohistochemistry (IHC) labeling using rabbit monoclonal antibody to
SLC7A5 (1:300 dilution; Epitomics, Burlingame, CA, USA) for 1 h at
room temperature. Envision (biotin-free) peroxidase-based detection
kit (Dako, Glostrup, Denmark) for mouse/rabbit antibodies was then
used with the red AEC or brown DAB substrate-chromogen (Vector
Labs, Burlingame, CA) followedby hematoxylin nuclear counterstain-
ing as previ ously described [25,26]. Alternatively, we performeddou-
ble immunofluorescent (IF) staining[25] where SLC7A5 antibody was
visualized with the use of horse-radish peroxidase (HRP)-coupled
anti-rabbit IgG(Fab)
2
and tyramide-FITC for green fluorescence
followed by mAB to TTF1 and biotinylated secondary antibody treat-
ments (all from Dako), developed with streptavidin-texas red for red
fluorescence (Vector). Specificity of the IHC-staining was determined
by negativecontrol staining, using non-immune control rabbit serum
(Dako) in place of the primary antibody (data not shown). Finally,
tissue-reactivities for the antibodies (percentage of positive cells)
were evaluated for each case, as described earlier [26],byusinga
3D-Histotech-Zeiss slide-scanner and Mirax-viewer software
program (3D-Histotech, Budapest, Hungary).
3. Results
3.1. Overexpression of miR-126 inhibits proliferation of SCLC cells by
causing delay in the G1 phase
To understand the role of miR-126 in the proliferation of SCLC
cells, H69 cells were transiently transfected with miR-126
Fig. 2. miR-126 suppresses SLC7A5 protein production. H69 cells were transfected with pre-miR-126 or NegmiR (negative control). Cell lysates and total RNA were prepared
at 96 h post-transfection. (A) Representative western blot for SLC7A5 and PLK2 in transfected H69 cells. GAPDH protein levels were used for normalization. The numbers
above the blot indicate normalized protein amounts relative to the negative control, as determined by densitometry. (B) Overexpression of mature miR-126 in transfected
cells determined by semi-quantitative RT-PCR. PCR products were visualized after electrophoresis in an EtBr-stained agarose gel. (C) SCL7A5, but not PLK2 protein production
is suppressed by miR-126, as determined by western analyses. (D) Effect of miR-126 overexpression on SLC7A5 and PLK2 mRNA levels, as determined by qRT-PCR.
E. Miko et al. / FEBS Letters 585 (2011) 1191–1196 1193
precursor, or the negative control miRNA, and cell numbers were
monitored for 96 h. As expected, transfection of pre-miR-126 into
H69 cells resulted in increased miR-126 expression compared to
non-transfected or NegMiR control-transfected cells (Fig. 2B). Over-
expression of miR-126 resulted in a significantly decreased prolifer-
ation of H69 cells, evident from 72 h post-transfection (Fig. 1A).
Similar observations were made for another SCLC cell line, HTB-
172 (Supplementary Fig. 1) On the other hand, overexpression of
miR-199a, which is also down-regulated in H69 cells, had no effect
on H69 cell proliferation (data not shown). Flow cytometric cell
cycle analysis at two time points (72 and 96 h post-transfection)
revealed an increasing percentage of miR-126-transfected cells in
the G1 phase over time, and a concomitant decrease in the percent-
age of cells in the G2/M phase. (Fig. 1B and C).
3.2. Overexpression of miR-126 suppresses SLC7A5 expression at both
the RNA and the protein level
To identify potential targets for miR-126 that might play a role
in regulating proliferation of SCLC cells, we first performed an in
silico analysis using the miRNA target prediction databases Target-
Scan and PicTar (Table 1). However, with the exception of SLC7A5,
none of the validated or doubly-predicted target genes are known
to be overexpressed in SCLC cell lines or tumors [11]. SLC7A5 pro-
tein overexpression in SCLC is in accordance with the previously
described downregulation of miR-126 expression. Therefore, we
selected SLC7A5 for further studies to analyze the role of
miR-126 in the cell cycle regulation of SCLC.
Table 1
Predicted and validated targets for miR-126.
Gene symbol Enscmbl ID Target
scan
PicTar Experimentally validated
CRK NM 016823 + + +
PLK2 NM 006622 + + +
SLC7A5 NM 003486 + +
PTPN9 NM 002833 + +
FBX033 NM 203301 + +
RGS3 NM 021106 + +
SPRED1 NM 152594 + +
TOM1 NM 005488 + +
IRS1 NM 005544 + +
HOXA9 NM 002142 + +
VCAM1 NM 001078 +
PIK3R2 NM 005027 +
SOX2 NM 003106 +
Fig. 3. Suppression of SLC7A5 production by RNAi delays H69 cells in the G1 phase. H69 cells were transfected with siRNAs specific to SLC7A5 (siSLC7A5) or PLK2 (siPLK2), or
with the negative control siRNA. (A) Representative western blot for SLC7A5 and PLK2 in siRNA-transfected H69 cells. GAPDH protein levels were used for normalization. The
numbers above the blot indicate normalized protein amounts relative to the negative control, as determined by densitometry. (B) Representative cell cycle analysis of H69
cells at 96 h post-transfection, analyzed on a FACSArray bioanalyzer. (C) Suppression of SLC7A5 production by RNAi delays H69 cells in the G1 phase. Cells were analyzed on a
FACSArray bioanalyzer at 72 or 96 h post-transfection. Asterisks indicate significant t-test results (P< 0.05).
1194 E. Miko et al. / FEBS Letters 585 (2011) 1191–1196
We next investigated the effect of miR-126 overexpression on
SLC7A5 and PLK2 expression. miR-126 overexpression in H69 cells
caused more than a 50% reduction in SLC7A5 mRNA levels, and also
a slight suppression of PLK2 mRNA expression, as determined by
qRT-PCR (Fig. 2C). Subsequent western blot analysis of SLC7A5 and
PLK2 demonstrated that while miR-126 overexpression resulted in
decreased SLC7A5 protein levels, PLK2 protein levels did not change
significantly (Fig. 2A and B). SLC7A5 expression was also suppressed
in pre-miR-126 transfected HTB-172 cells (Supplementary Fig. 2).
3.3. Suppression of SLC7A5 by RNA interference (RNAi) delays SCLC
cells in the G1 phase
To better understand the effect of SLC7A5 in SCLC cell cycle con-
trol, we utilized RNAi to specifically suppress SLC7A5 production in
H69 cells, and performed cell cycle analysis by flow cytometry at
72 and 96 h post-transfection. Transfection of specific siRNA into
H69 cells resulted in significantly lower SLC7A5 expression when
compared to the negative control siRNA (Fig. 3A). Similarly to the
effect of miR-126 overexpression, suppression of SLC7A5 resulted
in an increasing percentage of transfected cells in the G1 phase
over time, and a concomitant decrease in the percentage of cells
in the G2/M phase, when compared to the negative control siRNA
(Fig. 3B and C). In contrast, specific suppression of PLK2 expression
by RNAi had no such effect on the cell cycle distribution of trans-
fected H69 cells.
3.4. SLC7A5 is a direct target of miR-126
To confirm that SLC7A5 is a molecular target of miR-126, as sug-
gested by the previous experiments, we constructed a luciferase
reporter vector containing 331 bp of the SLC7A5 3
0
UTR, including
the predicted miR-126 binding site (WT-UTR). We also constructed
a control luciferase vector with the miR-126 binding site deleted
from the SLC7A5 3
0
UTR (DEL-UTR) (Fig. 4A). The sequenced
plasmids showed 100% identity with the SLC7A5 3
0
UTR, and the
intended deletion in the control vector (data not shown). H69 cells
were transiently transfected with the WT-UTR-luciferase or the
DEL-UTR-luciferase vector and with pre-miR-126. Co-transfection
of WT-UTR with pre-miR-126 resulted in a significant decrease in
luciferase protein levels; however, deletion of the miR-126 binding
site from the SLC7A5 3
0
UTR abolished this effect of miR-126
(Fig. 4B).
The correlation of miR-126 and SLC7A5 expression was also
investigated in 12 primary SCLC tumor samples, using immunohis-
tochemistry (IHC) with SLC7A5-specific antibody. SLC7A5 expres-
sion was not detectable in normal lung tissue, which is in
accordance with previous observations (Supplementary Fig. 3A, in-
set, and [12]). In contrast, the SCLC tumors tested positive for
SLC7A5 protein expression in fact, 8 tumors contained more than
70% SLC7A5-positive cells. As demonstrated by the double IF
stained specimens, the majority of tumor cells exhibited nuclear
staining for TTF1 (typical feature for SCLC) with SLC7A5 co-expres-
sion (Supplementary Fig. 3B). The tumor samples analyzed with
IHC were the same samples analyzed before for aberrant miR-
126 expression [8]. Since all 12 SCLC tumors overexpressed SLC7A5
and under-expressed miR-126, the inverse correlation between the
expression levels of miR-126 and its target could be corroborated
in primary tumors (Fig. 5).
4. Discussion
In the present work we demonstrate for the first time that
miR-126 overexpression has a negative effect on SCLC cell prolifer-
ation, by delaying cells in the G1 phase of the cell cycle. However,
miR-126 is not just an anti-proliferative miRNA; rather, it appears
to have multiple functions depending on the cell type and the
actual cellular environment. This is underscored by the observa-
tions, that not all validated miR-126 target mRNAs are affected
by miR-126 in every cell type. Interestingly, in SCLC cells
miR-126 overexpression does not suppress PLK2 expression, even
though PLK2 was shown to be a bona fide target for miR-126 in
CBF acute myeloid leukemia (AML) cells. A similar observation
can be made for TOM1, which is targeted by miR-126 in CF airway
epithelium cells, but not in MCF7 cells [4,12]; or for SPRED1, which
is targeted in HUVEC cells, but not in AML cell lines [5,13].Itis
presently unclear how certain target mRNAs are presented to,
and others are protected from miR-126 in these experimental
setups, but the resulting target selectivity could contribute to the
varied functions of miR-126.
Fig. 4. SLC7A5 is a direct target of miR-126. (A) The predicted miR-126 binding site
in the wild type SLC7A5 3
0
UTR (WT-UTR), and in the deleted construct (DEL-UTR).
(B) Relative luciferase activity of the SLC7A5 WT-UTR and the DEL-UTR luciferase
constructs in H69 cells transfected with miR-126 or the negative control (NegmiR).
Fig. 5. SLC7A5 and miR-126 expression levels are inversely correlated in primary
SCLC tumors. Relative expression levels of mature miR-126 (left Yaxis) were
determined in primary SCLC tumor specimens by qRT-PCR [8]. Overexpression of
SLC7A5 in the same panel of primary SCLC tumors was determined by immuno-
histochemistry, using an SLC7A5-specific monoclonal antibody. Percentage of
SLC7A5-positive neoplastic cells was determined for each tumor specimen (right
Yaxis). Normal lung tissue exhibited no SLC7A5-specific staining (Supplementary
Fig. 3).
E. Miko et al. / FEBS Letters 585 (2011) 1191–1196 1195
Importantly, we identified a novel target of miR-126 in SCLC
cells, SLC7A5, which is the light chain of the heterodimeric 4F2
amino acid transporter. SLC7A5 is overexpressed in many cancer
types, including SCLC, and its expression levels are usually corre-
lated to cancer progression and aggressiveness [14–16].We
demonstrated that in SCLC cells, similarly to other tumor types
[17–19], suppression of SLC7A5 expression has an anti-proliferative
effect. SCL7A5 suppression or miR-126 overexpression both delay
SCLC cells in the G1 phase, suggesting that the effect of miR-126
on the cell cycle is at least in part mediated through SLC7A5. Con-
sequently, decreased miR-126 expression contributes to high
SLC7A5 expression in SCLC cells, and, thus, may ensure efficient
transport of essential amino acids in the rapidly proliferating
tumor cells.
On the other hand, miR-126 may also be involved in a more
direct regulation of the cell cycle in SCLC. Glutamine–leucine
exchange by SLC7A5 was shown to activate the nutrient and
growth factor integrating kinase mTOR, which in turn phosphory-
lates p70S6 kinase 1 and 4EBP1, leading to the production of
growth promoting proteins [20]. Removal of miR-126 from the
regulatory network may enhance the existing positive feedback
between SLC7A5 and mTOR [21,22], and can contribute signifi-
cantly to the proliferative potential of the tumor cells. In addition,
miR-126 is capable of targeting the PI3K/Akt pathway as well,
either through PIK3R2 or some other mechanisms [1,2,5]. It should
be noted that both the PI3K/Akt and the mTOR pathways are
indeed active in a large percentage of SCLC tumors [23–26].
In summary, our work has identified miR-126 as an important
negative regulator of the growth and proliferation of SCLC cells,
which probably fine-tunes the activity of the PI3K/Akt/mTOR
network through multiple targets, including SLC7A5. However,
miR-126 may have additional functions in the tumor stroma: in
normal endothelial cells it is a positive regulator of angiogenesis,
and it may have interesting functions in regulating the immune
response. Therefore, more research is needed to understand the
complex role of miR-126 in the growth, survival and progression
of SCLC tumors in vivo, and to determine how miR-126 may poten-
tially be exploited as an anti-tumor agent.
Acknowledgments
This work was supported by grants from the National Office for
Research and Technology (B.S.: NKFP 2004 OM-00427 and GVOP-
3.1.1.-2004-05-0263/3.0; Z.B.: GVOP-3.2.1-2004-04-0351/3.0;
B.D.: NKTH-TECH-08-A1-2008-0228; Á.L.: TÁMOP 4.2.1./B-09/1/
KONV-2010-0007), by grants from the Hungarian Research Foun-
dation (Z.B.: OTKA T046945; Á.L.: OTKA 81676), and by grants from
the Hungarian Scientific Research Fund (Z.B.: OMFB-01626/2006).
Appendix A. Supplementary data
Supplementary data associated with this article can be found, in
the online version, at doi:10.1016/j.febslet.2011.03.039.
References
[1] Guo, C., Sah, J.F., Beard, L., Willson, J.K., Markowitz, S.D. and Guda, K. (2008)
The noncoding RNA, miR-126, suppresses the growth of neoplastic cells by
targeting phosphatidylinositol 3-kinase signaling and is frequently lost in
colon cancers. Genes Chromosom. Cancer 47, 939–946.
[2] Wang, X.-C. et al. (2011) Expression and function of miRNA in postoperative
radiotherapy sensitive and resistant patients of non-small cell lung cancer.
Lung Cancer 72, 92–99.
[3] Zhong, M., Ma, X., Sun, C. and Chen, L. (2010) MicroRNAs reduce tumor growth
and contribute to enhance cytotoxicity induced by gefitinib in non-small cell
lung cancer. Chem. Biol. Interact. 184, 431–438.
[4] Zhang, J., Du, Y.Y., Lin, Y.F., Chen, Y.T., Yang, L., Wang, H.J. and Ma, D. (2008) The
cell growth suppressor, mir-126, targets IRS-1. Biochem. Biophys. Res.
Commun. 377, 136–140.
[5] Fish, J.E. et al. (2008) MiR-126 regulates angiogenic signaling and vascular
integrity. Dev. Cell 15, 272–284.
[6] Wang, S. et al. (2008) The endothelial-specific microRNA miR-126 governs
vascular integrity and angiogenesis. Dev. Cell 15, 261–271.
[7] Barshack, I. et al. (2010) MicroRNA expression differentiates between primary
lung tumors and metastases to the lung. Pathol. Res. Pract. 206, 578–584.
[8] Feng, R. et al. (2010) miR-126 functions as a tumour suppressor in human
gastric cancer. Cancer Lett. 298, 50–63.
[9] Li, X., Shen, Y., Ichikawa, H., Antes, T. and Goldberg, G.S. (2009) Regulation of
miRNA expression by Src and contact normalization: effects on nonanchored
cell growth and migration. Oncogene 28, 4272–4283.
[10] Miko, E., Czimmerer, Z., Csanky, E., Boros, G., Buslig, J., Dezso, B. and Scholtz, B.
(2009) Differentially expressed microRNAs in small cell lung cancer. Exp. Lung
Res. 35, 646–664.
[11] Kaira, K. et al. (2008) Expression of L-type amino acid transporter 1 (LAT1) in
neuroendocrine tumors of the lung. Pathol. Res. Pract. 204, 553–561.
[12] Oglesby, I.K., Bray, I.M., Chotirmall, S.H., Stallings, R.L., O‘Neill, S.J., McElvaney,
N.G. and Greene, C.M. (2010) miR-126 is downregulated in cystic fibrosis
airway epithelial cells and regulates TOM1 expression. J. Immunol. 184, 1702–
1709.
[13] Li, Z. et al. (2008) Distinct microRNA expression profiles in acute myeloid
leukemia with common translocations. Proc. Natl. Acad. Sci. USA 105, 15535–
15540.
[14] Kaira, K. et al. (2009) CD98 expression is associated with poor prognosis in
resected non-small-cell lung cancer with lymph node metastases. Ann. Surg.
Oncol. 16, 3473–3481.
[15] Nakanishi, K. et al. (2007) Expression of LAT1 predicts risk of progression of
transitional cell carcinoma of the upper urinary tract. Virchows Arch. 451,
681–690.
[16] Sakata, T. et al. (2009) L-type amino-acid transporter 1 as a novel biomarker
for high-grade malignancy in prostate cancer. Pathol. Int. 59, 7–18.
[17] Fan, X., Ross, D.D., Arakawa, H., Ganapathy, V., Tamai, I. and Nakanishi, T.
(2010) Impact of system L amino acid transporter 1 (LAT1) on proliferation of
human ovarian cancer cells: a possible target for combination therapy with
anti-proliferative aminopeptidase inhibitors. Biochem. Pharmacol. 80, 811–
818.
[18] Nawashiro, H. et al. (2006) L-type amino acid transporter 1 as a potential
molecular target in human astrocytic tumors. Int. J. Cancer 119, 484–492.
[19] Shennan, D.B. and Thomson, J. (2008) Inhibition of system L (LAT1/CD98hc)
reduces the growth of cultured human breast cancer cells. Oncol. Rep. 20,
885–889.
[20] Nicklin, P. et al. (2009) Bidirectional transport of amino acids regulates mTOR
and autophagy. Cell 136, 521–534.
[21] Edinger, A.L. and Thompson, C.B. (2002) Akt maintains cell size and survival by
increasing mTOR-dependent nutrient uptake. Mol. Biol. Cell 13, 2276–2288.
[22] Liu, X.M., Reyna, S.V., Ensenat, D., Peyton, K.J., Wang, H., Schafer, A.I. and
Durante, W. (2004) Platelet-derived growth factor stimulates LAT1 gene
expression in vascular smooth muscle: role in cell growth. FASEB J. 18, 768–
770.
[23] Blackhall, F.H., Pintilie, M., Michael, M., Leighl, N., Feld, R., Tsao, M.S. and
Shepherd, F.A. (2003) Expression and prognostic significance of kit, protein
kinase B, and mitogen-activated protein kinase in patients with small cell lung
cancer. Clin. Cancer Res. 9, 2241–2247.
[24] Moore, S.M., Rintoul, R.C., Walker, T.R., Chilvers, E.R., Haslett, C. and Sethi, T.
(1998) The presence of a constitutively active phosphoinositide 3-kinase in
small cell lung cancer cells mediates anchorage-independent proliferation via
a protein kinase B and p70s6k-dependent pathway. Cancer Res. 58, 5239–
5247.
[25] Pardo, O.E., Arcaro, A., Salerno, G., Tetley, T.D., Valovka, T., Gout, I. and Seckl,
M.J. (2001) Novel cross talk between MEK and S6K2 in FGF-2 induced
proliferation of SCLC cells. Oncogene 20, 7658–7667.
[26] Razzini, G., Berrie, C.P., Vignati, S., Broggini, M., Mascetta, G., Brancaccio, A. and
Falasca, M. (2000) Novel functional PI 3-kinase antagonists inhibit cell growth
and tumorigenicity in human cancer cell lines. FASEB J. 14, 1179–1187.
1196 E. Miko et al. / FEBS Letters 585 (2011) 1191–1196
... SLC7A5 is known to be highly expressed in many cancers [21,22]. In KRASmutant colorectal cancer cells, SLC7A5 maintained intracellular amino acid level and activated mTOR pathway, which finally supported cell proliferation [23]; in small cell lung cancer, increased SLC7A5 promoted cell growth [24]. However, the current research on SLC7A5 in GC is insufficient. ...
Article
Full-text available
Gastric cancer (GC) is a common malignant tumor worldwide, especially in East Asia, with high incidence and mortality rate. Epigenetic modifications have been reported to participate in the progression of gastric cancer, among which m⁶A is the most abundant and important chemical modification in RNAs. Fat mass and obesity-associated protein (FTO) is the first identified RNA demethylase but little is known about its role in gastric cancer. In our study, data from TCGA and clinical samples showed that FTO was highly expressed in gastric cancer tissues. Kaplan–Meier plotter suggested that patients with the high level of FTO had a poor prognosis. In vitro and in vivo experiments confirmed the role of FTO in promoting gastric cancer cell proliferation. Mechanistically, we found that FTO bound to circFAM192A at the specific site and removed the m⁶A modification in circFAM192A, protecting it from degradation. CircFAM192A subsequently interacted with the leucine transporter solute carrier family 7 member 5 (SLC7A5) and enhancing its stability. As a result, an increased amount of SLC7A5 was on the membrane, which facilitated leucine uptake and activated the mTOR signaling pathway. Therefore, our study demonstrated that FTO promoted gastric cancer proliferation through the circFAM192A/SLC7A5 axis in the m⁶A-dependent manner. Our study shed new light on the role of FTO in gastric cancer progression. Supplementary Information The online version contains supplementary material available at 10.1186/s43556-024-00172-4.
... Cells were incubated with the transfection complexes for 6h before replacing the medium. Cells were replaced with fresh growth medium daily (23). ...
... The upregulation of SLC7A5 has been observed in several types of cancers, including lung and gastric cancers. Its expression has been shown to be associated with cancer cell proliferation, migration, invasion, and metastasis [79,80]. In addition, it has been demonstrated that SLC7A5 is upregulated in RB tissue [81]. ...
Article
Retinoblastoma (RB) is a common cancer in infants and children. It is a curable disease; however, a delayed diagnosis or treatment makes the treatment difficult. Genetic mutations have a central role in the pathogenesis of RB. Genetic materials such as RNAs (coding and non-coding RNAs) are also involved in the progression of the tumor. Circular RNA (circRNA) is the most recently identified RNA and is involved in regulating gene expression mainly through "microRNA sponges". The dysregulation of circRNAs has been observed in several diseases and tumors. Also, various studies have shown that circRNAs expression is changed in RB tissues. Due to their role in the pathogenesis of the disease, circRNAs might be helpful as a diagnostic or prognostic biomarker in patients with RB. In addition, circRNAs could be a suitable therapeutic target to treat RB in a targeted therapy approach.
... The role of miR-126 in modulating sensitivity to therapy has been noted with altered sensitivity to cisplatin in HCC through the repression of insulin receptor substrate 1 expression (55), or to sorafenib through the downregulation of sprout-related EVH1 domain-containing protein 1 (56). MiR-126-3p has also been indicated to elicit tumor suppressor effects in other gastrointestinal and other solid tumor types, such as lung cancer (57,58). ...
Article
Full-text available
Extracellular vesicles (EVs) and their contents are gaining recognition as important mediators of intercellular communication through the transfer of bioactive molecules, such as non‑coding RNA. The present study comprehensively assessed the microRNA (miRNA/miR) content within EVs released from HepG2 liver cancer (LC) cells and LX2 hepatic stellate cells (HSCs) and determined the contribution of EV miRNA to intercellular communication. Using both transwell and spheroid co‑cultures of LC cells and HSCs, miR‑126‑3p within EV was established as a mediator of HSC to LC cell communication that influenced tumor cell migration and invasion, as well as the growth of multicellular LC/HSC spheroids. Manipulation of miR‑126‑3p either by enforced expression using pre‑miR‑126‑3p or by inhibition using antimiR‑126‑3p did not alter tumor cell viability, proliferation or sensitivity to either sorafenib or regorafenib. By contrast, enforced expression of miR‑126‑3p decreased tumor‑cell migration. Knockdown of miR‑126‑3p in tumor cells increased disintegrin and metalloproteinase domain‑containing protein 9 (ADAM9) expression and in HSCs increased collagen‑1A1 accumulation with an increase in compactness of multicellular spheroids. Within LC/HSC spheroids, ADAM9 and vascular endothelial growth factor expression was increased by silencing of miR‑126‑3p but diminished with the restoration of miR‑126‑3p. These studies implicate miR‑126‑3p in functional effects on migration, invasion and spheroid growth of tumor cells in the presence of HSCs, and thereby demonstrate functional EV‑RNA‑based intercellular signaling between HSCs and LC cells that is directly relevant to tumor‑cell behavior.
Article
Radiotherapy resistance remains a huge hindrance in treating non-small cell lung carcinoma (NSCLC). Hyperthermia, a treatment method that raises the cell temperature to treat tumors, has been putting in clinical application combined with conventional chemotherapy and radiotherapy to enhance their effectiveness in NSCLC treatment. However, the specific mechanism of this combination therapy has not been extensively researched. In this study, we established a radiation-resistant NSCLC cell line by sequential radiation exposure. It was shown that the combination of hyperthermia and radiotherapy suppressed NSCLC xenograft tumor growth and increased radiation-induced apoptosis. Furthermore, downregulation of miR-126-5p was found in radio-resistant NSCLC cells. The results of bioinformatics analysis based on ENCORI showed that RAD50 is a direct target of miR-126-5p. Overexpressed miR-126-5p increased radiation-induced cell death by suppressing RAD50 expression. Hyperthermia treatment increases miR-126-5p and decreases RAD50, leading to more unrepaired DNA damage and greater cellular death. In conclusion, hyperthermia enhances NSCLC cells’ radio-sensitivity via miR-126-5p/RAD50 axis.
Article
Full-text available
For several decades, most lung cancer investigations have focused on the search for mutations in candidate genes; however, in the last decade, due to the fact that most of the human genome is occupied by sequences that do not code for proteins, much attention has been paid to non-coding RNAs (ncRNAs) that perform regulatory functions. In this review, we principally focused on recent studies of the function, regulatory mechanisms, and therapeutic potential of ncRNAs including microRNA (miRNA), long ncRNA (lncRNA), and circular RNA (circRNA) in different types of lung cancer.
Article
Full-text available
Carcinoma of the lung is among the most common types of cancer globally. Concerning its histology, it is categorized as a non-small cell carcinoma (NSCLC) and a small cell cancer (SCLC) subtype. MicroRNAs (miRNAs) are a member of non-coding RNA whose nucleotides range from 19 to 25. They are known to be critical regulators of cancer via epigenetic control of oncogenes expression and by regulating tumor suppressor genes. miRNAs have an essential function in a tumorous microenvironment via modulating cancer cell growth, metastasis, angiogenesis, metabolism, and apoptosis. Moreover, a wide range of information produced via several investigations indicates their tumor-suppressing, oncogenic, diagnostic assessment, and predictive marker functions in different types of lung malignancy. miRNA mimics or anti-miRNAs can be transferred into a lung cancer cell, with possible curative implications. As a result, miRNAs hold promise as targets for lung cancer treatment and detection. In this study, we investigate the different functions of various miRNAs in different types of lung malignancy, which have been achieved in recent years that show the lung cancer-associated regulation of miRNAs expression, concerning their function in lung cancer beginning, development, and resistance to chemotherapy, also the probability to utilize miRNAs as predictive biomarkers for therapy reaction.
Article
Full-text available
Cystic fibrosis (CF) is one of the most common lethal genetic diseases in which the role of microRNAs has yet to be explored. Predicted to be regulated by miR-126, TOM1 (target of Myb1) has been shown to interact with Toll-interacting protein, forming a complex to regulate endosomal trafficking of ubiquitinated proteins. TOM1 has also been proposed as a negative regulator of IL-1beta and TNF-alpha-induced signaling pathways. MiR-126 is highly expressed in the lung, and we now show for the first time differential expression of miR-126 in CF versus non-CF airway epithelial cells both in vitro and in vivo. MiR-126 downregulation in CF bronchial epithelial cells correlated with a significant upregulation of TOM1 mRNA, both in vitro and in vivo when compared with their non-CF counterparts. Introduction of synthetic pre-miR-126 inhibited luciferase activity in a reporter system containing the full length 3'-untranslated region of TOM1 and resulted in decreased TOM1 protein production in CF bronchial epithelial cells. Following stimulation with LPS or IL-1beta, overexpression of TOM1 was found to downregulate NF-kappaB luciferase activity. Conversely, TOM1 knockdown resulted in a significant increase in NF-kappaB regulated IL-8 secretion. These data show that miR-126 is differentially regulated in CF versus non-CF airway epithelial cells and that TOM1 is a miR-126 target that may have an important role in regulating innate immune responses in the CF lung. To our knowledge, this study is the first to report of a role for TOM1 in the TLR2/4 signaling pathways and the first to describe microRNA involvement in CF.
Article
Full-text available
The purpose of this study was to evaluate the prognostic value of L-type amino acid transporter 1 (LAT1) and 4F2 heavy chain (CD98) expression in resectable non-small-cell lung cancer (NSCLC) patients with N1 and N2 nodal involvement. A total of 220 consecutive patients were retrospectively reviewed. Immunohistochemical expression of LAT1, CD98, Ki-67 labeling index, vascular endothelial growth factor (VEGF), and microvessel density (MVD) was correlated with clinical features and prognosis of patients after complete resection of the tumor. Positive expression of LAT1 and CD98 was recognized in 60% (132/220) and 47% (103/220), respectively (P = 0.021). A positive rate of LAT1 expression was significantly higher in squamous cell carcinoma (SQC) (91%; 65/71) and large cell carcinoma (LCC) (82%; 9/11) than in adenocarcinoma (AC) (42%; 58/138). Moreover, a positive rate of CD98 expression was also significantly higher in SQC (76%; 54/71) and LCC (73%; 8/11) than in AC (30%; 42/138). LAT1 expression was significantly correlated with CD98, Ki-67 labeling index, VEGF, and MVD. The 5-year survival rates of LAT1-positive and LAT1-negative patients and CD98-positive and CD98-negative patients, were 43% and 48% (P = 0.1043), respectively and 39% and 50% (P = 0.0239), respectively. Multivariate analysis confirmed that positive expression of CD98 was an independent factor for predicting a poor prognosis. In our limited series, CD98 is a pathological factor that predicts prognosis in resectable adenocarcinoma patients with N2 disease.
Article
Full-text available
Transformation by the Src tyrosine kinase (Src) promotes nonanchored cell growth and migration. However, nontransformed cells can force Src-transformed cells to assume a normal morphology and phenotype by a process called 'contact normalization'. It has become clear that microRNA (miRNA) can affect tumorigenesis by targeting gene products that direct cell growth and migration. However, the roles of miRNA in Src transformation or contact normalization have not yet been reported. We examined the expression of 95 miRNAs and found 9 of them significantly affected by Src. In this study, we report that miR-218 and miR-224 were most significantly induced by Src, but not affected by contact normalization. In contrast, miR-126 was most significantly suppressed by Src and was induced by contact normalization in transformed cells. Mir-126 targets Crk, a component of the focal adhesion network that participates in events required for tumor cell migration. Accordingly, we show that miR-126 expression correlates inversely with Crk levels, motility and the invasive potential of human mammary carcinoma cells. Moreover, we show that miR-224 expression promotes nonanchored growth of nontransformed cells. These data reveal novel insights into how Src regulates miRNA expression to promote hallmarks of tumor cell growth and invasion, and how nontransformed cells can affect miRNA expression in adjacent tumor cells to inhibit this process.
Article
To investigate the different miRNA expression profiles of postoperative radiotherapy sensitive and resistant patients of non-small cell lung cancer, explore their potential role and find some radio-sensitivity markers. Thirty non-small cell lung cancer patients who have been treated by postoperative radiotherapy were selected and were divided into radiotherapy sensitive group and resistant group according to overall survival and local or distant recurrence rate. Expression profile of miRNA in these two groups was detected by a microarray assay and the results were validated by quantitative RT-PCR and Northern blot. At the molecular level, the effect of one differently expressed miRNA (miR-126) on the growth and apoptosis of SK-MES-1 cells induced by irradiation was examined. Comparing with resistant patients, five miRNAs (miRNA-126, miRNA-let-7a, miRNA-495, miRNA-451 and miRNA-128b) were significantly upregulated and seven miRNAs (miRNA-130a, miRNA-106b, miRNA-19b, miRNA-22, miRNA-15b, miRNA-17-5p and miRNA-21) were greatly downregulated in radiotherapy sensitive group. Overexpression of miRNA-126 inhibited the growth of SK-MES-1 cells and promoted its apoptosis induced by irradiation. The expression level of p-Akt decreased in miRNA-126 overexpression group. After treating with phosphoinositidyl-3 kinase (PI3K) constitutively activator (IGF-1) and inhibitor (LY294002), miRNA-126 overexpression had no significant effects on the apoptosis of SK-MES-1 cells. We found 12 differently expressed miRNAs in the radiotherapy sensitive and resistant non-small cell lung cancer samples. Moreover, our results showed miRNA-126 promoted non-small cell lung cancer cells apoptosis induced by irradiation through the PI3K-Akt pathway.
Article
MicroRNAs have emerged as important gene regulators and are recognised as key players in carcinogenesis. In the present study, we show that miR-126 was significantly down-regulated in gastric cancer tissues compared with matched normal tissues and was associated with clinicopathological features, including tumour size, lymph node metastasis, local invasion and tumour-node-metastasis (TNM) stage. Ectopic expression of miR-126 in SGC-7901 gastric cancer cells potently inhibited cell growth by inducing cell cycle arrest in G0/G1 phase, migration and invasion in vitro as well as tumorigenicity and metastasis in vivo. Mechanistically, we identified the adaptor protein Crk as a target of miR-126. Taken together, our results suggest that miR-126 may function as a tumour suppressor in gastric cancer, with Crk as a direct target.
Article
Amino acids activate nutrient signaling via the mammalian target of rapamycin (mTOR), we therefore evaluated the relationship between amino acid transporter gene expression and proliferation in human ovarian cancer cell lines. Expression of three cancer-associated amino acid transporter genes, LAT1, ASCT2 and SN2, was measured by qRT-PCR and Western blot. The effects of silencing the LAT1 gene and its inhibitor BCH on cell growth were evaluated by means of cell proliferation and colony formation assays. The system L amino acid transporter LAT1 was up-regulated in human ovarian cancer SKOV3, IGROV1, A2780, and OVCAR3 cells, compared to normal ovarian epithelial IOSE397 cells, whereas ASCT2 and SN2 were not. BCH reduced phosphorylation of p70S6K, a down-stream effector of mTOR, in SKOV3 and IGROV1 cells, and decreased their proliferation by 30% and 28%, respectively. Although proliferation of SKOV3 (S1) or IGROV1 (I10) cells was unaffected by LAT1-knockdown, plating efficiency in colony formation assays was significantly reduced in SKOV3(S1) and IGROV1(I10) cells to 21% and 52% of the respective plasmid transfected control cells, SKOV3(SC) and IGROV(IC), suggesting that LAT1 affects anchorage-independent cell proliferation. Finally, BCH caused 10.5- and 4.3-fold decrease in the IC(50) value of bestatin, an anti-proliferative aminopeptidase inhibitor, in IGROV1 and A2780 cells, respectively, suggesting that the combined therapy is synergistic. Our findings indicate that LAT1 expression is increased in human ovarian cancer cell lines; LAT1 may be a target for combination therapy with anti-proliferative aminopeptidase inhibitors to combat ovarian cancer.
Article
For surgical pathologists, distinguishing whether a pulmonary neoplasm is primary or metastatic can be challenging, and current biomarkers do not always aid lung tumor classification. The tissue-associated expression of microRNA likely explains the remarkable finding that many tumors can be classified based solely on their microRNA expression signature. Here we show that microRNAs can serve as biomarkers for lung tumor classification. Using microRNA microarray data generated from 76 formalin-fixed, paraffin-embedded (FFPE) samples of either primary lung cancer or metastatic tumors to the lung, we have identified a set of microRNAs expressed differentially between these two groups. This set includes hsa-miR-182, which was most strongly over-expressed in the lung primary tumors, and hsa-miR-126, which was over-expressed in the metastatic tumors. The differential expression of this set of microRNAs was confirmed using qRT-PCR on a set of 54 samples. In light of our data, microRNA expression should be considered as a potential clinical biomarker for surgical pathologists faced with discerning the tumor type of an inscrutable lung neoplasm.
Article
MicroRNAs (miRNAs) have emerged as key post-transcriptional regulators of gene expression, involved in diverse physiological and pathological processes. An oncogenic or tumor-suppressive miRNA may have potential as a therapeutic target to control cancers. Gefitinib is a tyrosine kinase inhibitor that targets epidermal growth factor receptor (EGFR). H460 and A549 cells with EGFR receptor-independent over-activation of protein kinase B (Akt) or extracellular signal-regulated kinases (ERK) are significantly resistant to gefitinib. The first aim of this study was to confirm a role for three miRNAs (let-7a, hsa-miR-126, and hsa-miR-145) in the inhibition of proliferation in non-small cell lung cancer (NSCLC) cells. A second aim was to evaluate three miRNAs for their abilities to overcome cellular resistance and enhance the gefitinib cytotoxicity. The expression of miRNAs was estimated by real-time quantitative reverse transcription polymerase chain reaction (qRT-PCR). Cell proliferation was examined by sulforhodamine B assay and tumor xenografts were measured in SCID/beige mice. The activation of Akt and ERK was observed by Western blotting. Forced expression of individual miRNA suppressed the growth of two cell lines and xenografts. The effect varied among different miRNAs and cells. Restoration of hsa-miR-126 more obviously inhibited cell growth than did restoration of hsa-miR-145 in both cells, and the suppressive effect was more significant in H460 xenografts than in A549 xenografts. Western blotting revealed that the inhibition of cell proliferation resulted from the inhibition of the activation of Akt and ERK. Moreover, forced expression of miRNAs contributed to enhanced cytotoxicity induced by gefitinib in lung cancer cells; especially in hsa-miR-126, the highest value of half max inhibitory (IC50) was increased sixfold. These findings confirm that tumor-suppressive miRNAs can inhibit the growth of NSCLC cells and enhance the targeted agents cytotoxicity, suggesting novel potential approaches to an improvement in chemotherapy.
Article
Expression of microRNAs (miRNAs) is characteristically altered in cancer, and they may play a role in cancer development and progression. The authors performed microarray and real-time quantitative reverse transcriptase-polymerase chain reaction (RT-PCR) analyses to determine the miRNA expression profile of primary small cell lung cancer. Here we show that at least 24 miRNAs are differentially expressed between normal lung and primary small cell lung cancer (SCLC) tumors. These include miR-301, miR-183/96/182, miR-126, and miR-223, which are microRNAs deregulated in other tumor types as well; and other miRNAs, such as miR-374 and miR-210, not previously reported in association with lung cancer. The aberrant miRNA profile of SCLC may offer new insights in the biology of this aggressive tumor, and could potentially provide novel diagnostic markers.
Article
Amino acids are required for activation of the mammalian target of rapamycin (mTOR) kinase which regulates protein translation, cell growth, and autophagy. Cell surface transporters that allow amino acids to enter the cell and signal to mTOR are unknown. We show that cellular uptake of L-glutamine and its subsequent rapid efflux in the presence of essential amino acids (EAA) is the rate-limiting step that activates mTOR. L-glutamine uptake is regulated by SLC1A5 and loss of SLC1A5 function inhibits cell growth and activates autophagy. The molecular basis for L-glutamine sensitivity is due to SLC7A5/SLC3A2, a bidirectional transporter that regulates the simultaneous efflux of L-glutamine out of cells and transport of L-leucine/EAA into cells. Certain tumor cell lines with high basal cellular levels of L-glutamine bypass the need for L-glutamine uptake and are primed for mTOR activation. Thus, L-glutamine flux regulates mTOR, translation and autophagy to coordinate cell growth and proliferation.