Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Journal of Pathology

J Pathol 2015; 237: 152–165 ORIGINAL PAPER


Published online 4 June 2015 in Wiley Online Library
(wileyonlinelibrary.com) DOI: 10.1002/path.4562

Functional screening identifies MCT4 as a key regulator of breast


cancer cell metabolism and survival
Franziska Baenke,1# Sébastien Dubuis,2 Charlene Brault,3 Britta Weigelt,4 Beatrice Dankworth,3 Beatrice Griffiths,1
Ming Jiang,5 Alan Mackay,6 Becky Saunders,5 Bradley Spencer-Dene,7 Susana Ros,1 Gordon Stamp,7 Jorge S
Reis-Filho,4 Michael Howell,5 Nicola Zamboni3 and Almut Schulze1,3,8*
1 Gene Expression Analysis Laboratory, Cancer Research UK London Research Institute, UK
2 Institute of Molecular Systems Biology, ETH Zurich, Switzerland
3 Department of Biochemistry and Molecular Biology, Theodor-Boveri-Institute, Biocentre Am Hubland, Würzburg, Germany
4 Department of Pathology, Memorial Sloan Kettering Cancer Center, New York, NY, USA
5 High Throughput Screening Facility, Cancer Research UK London Research Institute, UK
6
Divisions of Molecular Pathology and Cancer Therapeutics, Institute of Cancer Research, Sutton, Surrey, UK
7 Experimental Histopathology, Cancer Research UK London Research Institute, UK
8 Comprehensive Cancer Centre Mainfranken, Würzburg, Germany

*Correspondence to: A Schulze, Department of Biochemistry and Molecular Biology, Theodor-Boveri-Institute, Biocenre, Am Hubland, 97074
Würzburg, Germany. E-mail: almut.schulze@uni-wuerzburg.de
#
Current address: Molecular Oncology Laboratory, Cancer Research UK Manchester Institute, Wilmslow Road, Manchester M20 4BX, UK.

Abstract
Metabolic reprogramming in cancer enhances macromolecule biosynthesis and supports cell survival. Oncogenic
drivers affect metabolism by altering distinct metabolic processes and render cancer cells sensitive to perturbations
of the metabolic network. This study aimed to identify selective metabolic dependencies in breast cancer by
investigating 17 breast cancer cells lines representative of the genetic diversity of the disease. Using a functional
screen, we demonstrate here that monocarboxylate transporter 4 (MCT4) is an important regulator of breast cancer
cell survival. MCT4 supports pH maintenance, lactate secretion and non-oxidative glucose metabolism in breast
cancer cells. Moreover, MCT4 depletion caused an increased dependence of cancer cells on mitochondrial respiration
and glutamine metabolism. MCT4 depletion reduced the ability of breast cancer cells to grow in a three-dimensional
(3D) matrix or as multilayered spheroids. Moreover, MCT4 expression is regulated by the PI3K–Akt signalling
pathway and highly expressed in HER2-positive breast cancers. These results suggest that MCT4 is a potential
therapeutic target in defined breast cancer subtypes and reveal novel avenues for combination treatment.
Copyright © 2015 Pathological Society of Great Britain and Ireland. Published by John Wiley & Sons, Ltd.

Keywords: breast cancer; metabolism; MCT4; lactate; PI3K–Akt pathway; HER2

Received 12 August 2014; Revised 23 March 2015; Accepted 4 May 2015

No conflicts of interest were declared.

Introduction Breast cancer comprises a heterogeneous group of


tumours. Clinically, it is classified by expression of
Metabolic transformation in cancer supports macro- oestrogen receptor (ER), progesterone receptor (PR)
molecule biosynthesis and is crucial for cancer cell sur- and human epidermal growth factor receptor 2 (HER2).
vival and tumour formation [1,2]. Many solid tumours Comprehensive genomic and transcriptomic analyses
exhibit increased glucose uptake and lactate secre- revealed that breast cancers can be divided into several
tion, a feature known as the Warburg effect. Most subgroups with distinct molecular and clinical features
metabolic alterations observed in cancer are the conse- [7–12]. These analyses have revealed the extent of the
quence of oncogene activation or inactivation of tumour diversity of the constellation of somatic genetic aberra-
suppressors [3], but genetic alterations in metabolic tions found in breast cancer; only a few genes are recur-
enzymes have also been observed. Cancer metabolism rently mutated, many genes are altered in a small number
is a highly dynamic network that adapts to changes in of cases, and each breast cancer may harbour a distinct
environmental conditions. However, metabolic repro- repertoire of somatic genetic changes [10,11,13,14].
gramming produces selective dependencies in cancer One common feature within the complex genetic land-
cells [4] and metabolic enzymes may provide suitable scape of breast cancer is the activation of the PI3K–Akt
targets for the development of novel treatment strate- signalling pathway [9–11]. Akt is one of the main
gies [5,6]. drivers of anabolic metabolism and activates glucose
Copyright © 2015 Pathological Society of Great Britain and Ireland. J Pathol 2015; 237: 152–165
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com
MCT4 supports breast cancer cell survival 153

uptake, glycolysis and lipid synthesis [15]. In addi- phenylhydrazone (FCCP) and oligomycin were from
tion to genetic heterogeneity, breast cancers also show Sigma-Aldrich.
environmental complexity with regions of high and
low delivery of nutrients and oxygen. The presence of siRNA screen and data analysis
low-oxygen niches in solid tumours (hypoxia) is asso-
ciated with poor clinical outcome and can promote ded- Cells were reverse-transfected with 37.5 nM Dharmacon
ifferentiation, chemoresistance, enhanced migration and siRNA SMARTpools of 231 genes in 96-well plates,
metastasis formation [16,17]. Identifying the molecular using Lullaby reagent (Oz Biosciences). After 24 h, the
dependencies of hypoxic cancer cells is integral to suc- culture medium was topped up with fresh medium; 96 h
cessful treatment. post-transfection, cells were fixed with 80% ice-cold
ethanol, stained with 4′ ,6-diamidino-2-phenylindole
In this study, we sought to identify specific metabolic
(DAPI) and cell numbers quantified by Acumen
dependencies of genetically diverse breast cancer cell
Explorer (TPP Labtech) (see supplementary material).
lines under normoxic and hypoxic conditions. Using an
unbiased functional RNA interference (RNAi) screen,
we found that the proton-linked monocarboxylate Intracellular pH, apoptosis and ROS levels
transporter 4 (MCT4) is an important regulator of Intracellular pH [pH(i)] was determined with 10 mM
breast cancer cell survival in vitro and in vivo. We SNARF-4 F 5-(and-6)-carboxylic acid (Molecular
found that MCT4 supports metabolite transport, pH Probes). The cells were trypsinized, washed and
homeostasis and redox balance in breast cancer cells. incubated with SNARF-4 F for 30 min. Changes in
Detailed metabolic characterization of MCT4-depleted fluorescence were measured using a Fortessa cytometer
cells revealed an increased dependence on oxida- (BD Biosciences). Cell death was determined 96 h
tive metabolism and glutaminolysis. MCT4 depletion post-transfection, using the Apo-ONE assay (Promega).
blocked the ability of breast cancer cells to grow in 3D For ROS detection, cells were incubated with 5 μM
culture systems and form xenograft tumours. Moreover, CM-H2 DCFDA (Molecular Probes) for 30 min at 37 ∘ C,
we found that MCT4 is regulated by the PI3K–Akt trypsinized, washed twice with phosphate-buffered
signalling axis and over-expressed in human breast saline (PBS), stained with DAPI and analysed on a
cancer tissue. The results of this study confirm MCT4 LSRII-SORP flow cytometer (Becton Dickinson).
as a potential drug target in breast cancer and suggest
strategies for patient stratification and combination Glycolytic rate and respiration
treatment.
Glycolytic rate was determined by monitoring the con-
version of 5-3 H-glucose to 3 H2 O. Cells were incubated
Materials and methods in glucose-free DMEM for 30 min before the addition
of 10 mM glucose spiked with 10 mCi 5-3 H-glucose for
1 h at 37 ∘ C. The reaction was quenched with 0.2 N
Additional methods not included in this section are pro-
HCl and 3 H2 O was captured by diffusion. Respiration
vided in the supplement (see supplementary material).
experiments were performed in a 96-well format, using a
Seahorse Bioscience XF96 Extracellular Flux Analyser.
Cell culture, antibodies and reagents Cells were incubated in glucose- and glutamine-free
Cell lines were obtained from the American Type DMEM (phenol red-free) for 2 h. The oxygen consump-
Culture Collection (ATCC) and LRI Cell Services. tion rate (OCR) was measured after injection of 10 mM
Cells were cultured in Dulbecco’s modified Eagle’s glucose and normalized to total protein content.
medium (DMEM)/F12 (1:1), with 2 mM L-glutamine
and penicillin–streptomycin. The medium was sup- 3D cultures and spheroid assays
plemented with 10% fetal calf serum (FCS; Gibco)
Cells were transfected with siRNA and grown as a
for the cancer cell lines and 5% horse serum, 20 ng/ml
monolayer, ie in two-dimensional (2D) culture. After
epithelial growth factor (EGF), 5 μg/ml hydrocorti-
24 h, the cells were trypsinized and seeded in an on-top
sone, 10 μg/ml insulin and 100 ng/ml cholera toxin for
Matrigel assay [18]. For spheroid formation, cells were
the non-malignant cell lines. MCT4 antibodies were
mixed with 2% Matrigel in culture medium and placed in
obtained from Santa Cruz (cat. no. sc-50329), BSG from
96-well ultra-low attachment plates (Costar). Spheroid
Novus (cat. no. NBP1-19677) and MCT1 from Millipore formation was initiated by centrifugation at 650 × g and
(cat. no. AB3538P). Horseradish peroxidase-conjugated the cultures were incubated for 7–14 days. At least
anti-β-actin and secondary antibodies were from six spheroids/group were photographed and size was
Abcam and Sigma-Aldrich, respectively. All other calculated using Photoshop (CS5.1).
antibodies were from Cell Signaling. Hydrocortisone,
EGF, PI-103 and metformin were from Calbiochem;
PD98059, UO126 and LY294002 from New England Co-culture experiments
Biolabs; and lapatinib was from LC Laboratories. MDA-MB-468 cells were plated together with red flu-
Bis-2-(5-phenylacetamido-1,3,4-thiadiazol-2-yl)ethyl orescent protein-expressing murine cancer-associated
sulfide (BPTES), carbonyl cyanide 4-(trifluoromethoxy) fibroblasts (mCherry-CAFs; a gift from E. Sahai, CRUK
Copyright © 2015 Pathological Society of Great Britain and Ireland. J Pathol 2015; 237: 152–165
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com
154 F Baenke et al

LRI), at a 5:1 ratio, and reverse-transfected with siRNA is consistent with an induction of SLC16A3 expres-
targeting human MCT4 (37.5 nM Dharmacon siGenome sion in most breast cancer cell lines under hypoxia
MCT4 5). After 3 days, the cells were stained with (Figure 1D, E).
Annexin V–Pacific blue and propidium iodide (PI) and To confirm the differential sensitivity of breast cancer
apoptotic cells in the mCherry-negative population were cells towards SLC16A3 depletion, three representative
determined by FACS analysis. cell lines were chosen. Silencing of SLC16A3 caused
a substantial reduction in cell number in HCC1954
Xenograft experiments and MDA-MB-468 cells, while non-malignant
MCF10A cells were insensitive to SLC16A3 deple-
MDA-MB-468 cells (1.5 × 106 cells) expressing tion (Figure 1 F). Loss of viability in MDA-MB-468
inducible shRNA targeting MCT4 (Tet-PLKO-shMCT4 cells was further validated using four different siRNA
88) were mixed with 10% Matrigel and injected into sequences and efficient depletion of SLC16A3 mRNA
the mammary fat pad of nude mice (nu/nu). After was confirmed (see supplementary material, Figure
tumour initiation (day 10), the mice were divided into
S2A). Non-malignant MCF10A cells were insensitive
two cohorts (n = 5 each). One cohort was treated with
to SLC16A3 silencing despite efficient down-regulation
doxycycline (0.2 g/kg food pellet; Harlan D.98186)
of mRNA and protein by three of four siRNA sequences
and tumour growth was followed over 65 days. Animal
(see supplementary material, Figure S2B).
experiments were performed according to UK Home
Next, we investigated the effect of long-term silenc-
Office guidelines.
ing of SLC16A3 on viability of MCF10A cells and
MDA-MB-468 cells. Inducible shRNA expression
Statistical analysis resulted in efficient down-regulation of SLC16A3
Graphs were generated using GraphPad Prism 5.0 mRNA and protein expression in both cell lines (see
(GraphPad software). All experiments were performed supplementary material, Figure S2C, D), but only
independently at least twice, with at least three biologi- reduced the viability of MDA-MB-468 cells (see sup-
cal replicates; p values were generated using two-tailed plementary material, Figure S2C, D). Moreover, growth
unpaired Student’s t-test or Mann–Whitney U-test, as factors used for culture of MCF10A cells did not alter
indicated. the sensitivity of either cell line towards SCL16A3
ablation (see supplementary material, Figure S2E),
while expression of an RNAi-insensitive version of
SLC16A3 (MCT4) rescued the observed reduction in
Results
cell number caused by silencing of endogenous MCT4
(see supplementary material, Figure S2F).
Functional screen identifies metabolic enzymes Interestingly, meta-analysis of human gene expres-
required for breast cancer cell survival sion data revealed that SLC16A3 mRNA levels are
To identify metabolic liabilities in breast cancer, we per- elevated in invasive breast cancer compared to nor-
formed an unbiased RNA interference (RNAi) screen, mal breast tissue (Figure 1G). Increased expression of
targeting 231 enzymes, transporters and regulators SLC16A3 was also associated with tumour recurrence
involved in central carbon metabolism (Figure 1A). (Figure 1H).
The screen was performed under normoxic and hypoxic
conditions in 17 breast epithelial cell lines representa-
Expression MCT1 and MCT4 in breast cancer cells
tive of different disease subtypes (see supplementary
material, Figure S1). Cell numbers were determined SLC16A3 encodes the monocarboxylate transporter 4
after 4 days of silencing and used to calculate z-scores (MCT4), a lactate–proton symporter that regulates intra-
(see supplementary material, Table S1). For 12 genes cellular pH in cancer cells [19,20]. We investigated the
we confirmed the effect on cell number with at least expression levels of MCT4, the related monocarboxy-
two individual siRNA sequences (see supplementary late transporter 1 (MCT1) and the ancillary glycoprotein
material, Table S2). The gene with the strongest selec- basigin (BSG) across the cell line panel. Expression of
tivity for cancer cells was SLC16A3. Silencing of MCT4 and MCT1 mRNA was higher in non-malignant
SLC16A3 in normoxia caused a substantial reduction cells than in cancer cells, while BSG showed little vari-
in cell number in six of 14 cancer cell lines, while ation (Figure 2A). Interestingly, HER2+ breast cancer
not affecting non-malignant cells (Figure 1B). This cell lines showed higher levels of MCT4 mRNA expres-
was even more pronounced under hypoxic conditions, sion than the other subtypes (Figure 2B). Expression of
with eight of 14 cancer cell lines showing a substantial MCT4 and MCT1 protein was variable, with high lev-
response (Figure 1C; see also supplementary material, els of MCT1 protein found in some ER− HER2− cell
Table S1). A comparison of different disease subtypes lines (Figure 2C), consistent with a previous report [21].
revealed that the effect of SLC16A3 silencing was more However, ablation of MCT1 did not affect the viability
pronounced in HER2+ cells under normoxic conditions. of the breast cancer cell lines used here, despite efficient
Under hypoxic conditions, most breast cancer cell depletion of the protein (see supplementary material,
lines were sensitive to ablation of SLC16A3, which Table S1, Figure S2G).
Copyright © 2015 Pathological Society of Great Britain and Ireland. J Pathol 2015; 237: 152–165
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com
MCT4 supports breast cancer cell survival 155

A B Normoxia C Hypoxia
non- non-
cancer malign. cancer malign.

SLC16A3 SLC16A3
OXCT2 MT−ND4
ODC1 CHPT1
GPI GNPNAT1
GPX4 COX4I1
SOD1 ODC1
CHPT1 MT−ND2
OLAH PPAP2A
GPX2 IDH1
C12orf5 MT−ND6
HSD17B7 GGPS1
DLST PPAP2B
Glycolysis & galactose metabolism DDO RPE
CHKA OXCT2
TCA cycle
MVD GFPT2
Pentose phosphate pathway PGM3
Aminoacid metabolism LOC441996
PFKL
PHGDH ≤-2.0
Hexosamine metabolism GLUD2
PGAM4
Oxphos & ROS metabolism COX4I2
DHCR7 ACSS1
Fatty acid metabolism SREBF1 FDFT1
Cholesterol synthesis PDP2 PDP2
Metabolic regulation PMVK DHCR7
BT-20
BT-549
HCC1806
HCC1954
HCC38
HS-578T
MCF7
MDA-MB-231
MDA-MB-436
MDA-MB-468
SK-BR-3
T-47D
ZR-75-1
MCF10A
MCF12A
184B5
AU-565

Others MCAT
Total=231 siRNAs PMVK 0

BT-20
BT-549
HCC1806
HCC1954
HCC38
HS-578T
MCF7
MDA-MB-231
MDA-MB-436
MDA-MB-468
SK-BR-3
T-47D
ZR-75-1

184B5
MCF10A
MCF12A
AU-565
D Normoxia Hypoxia E
** ** ER+/HER2- ER-/HER2-
3 **** 3 *** HER2+ non-malign.
*** ****
2 2 20 20% O2
relative SLC16A3 mRNA
SLC16A3 z-score

0.5% O2
SLC16A3 z-score

1 1
15
0 0
-1 -1
10
-2 -2
-3 -3 5
-4 -4
-5 -5 0
2+

2+
n.
2-

2-
2-

2-
n.

184B5
BT-549
HCC1954

HCC1806
AU-565

HS-578T
BT-20
SK-BR-3

MCF10A
MCF12A
MDA-MB-468

MDA-MB-231
MDA-MB-436
ZR-75-1

T-47D

HCC38
MCF7
ig
ER

ER
ig

ER

ER
ER

ER
al
al

m
H

H
-/H

-/H
H

H
m

+/

+/
n-
n-

ER

ER
ER

ER
no
no

F HCC1954 MDA-MB-468 MCF10A


20 30 50 n.s.
*** ****
cell number (x1000)
cell number (x1000)

cell number (x1000)

40
15
20
**** ** 30
10 n.s.
20
10
5
10

0 0 0
M rl

M rl

M rl

M rl
T4

T4

T4

T4

T4
M rl

M rl
T4
si iCt

si iCt

si iCt

si iCt

si iCt

si iCt
C

C
C
s

20% O2 0.5% O2 20% O2 0.5% O2 20% O2 0.5% O2

G H
TCGA Zhao Richardson Finak
(p<0.001) (p<0.001) (p<0.001) (p<0.003)
2 1 6 2
SLC16A3 expression

SLC16A3 expression

SLC16A3 expression
SLC16A3 expression

0 4 1
0
-1 2 0
-2
-2 0 -1
-4 -3 -2 -2
-6 -4 -4 -3
a l
ar
a l

om ta

re yeance
st

om la
om ta
st

a l

ar e
ar

as
va nomcta
va nomas

om la

s
ye c
in c

at cur rs
ea

in u
ea

ca ive a

in c

3 ren
re
in u

rc du
rc lob
rc du
i e

3 rre
rc du

rc lob

br
br

rc br

at cu
al
al
al

ca e

ca ve
i
v

rm
m

re
rm

ca
ca
si

ca
s

si

no
va

no
no
no

in
in

in

Figure 1. Legend on next page.

Copyright © 2015 Pathological Society of Great Britain and Ireland. J Pathol 2015; 237: 152–165
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com
156 F Baenke et al

MCT4 supports pH regulation, redox balance to glucose, most likely due to impaired glycolytic flux
and non-oxidative glucose metabolism in breast (Figure 3E), indicating that MCT4 is required to main-
cancer cells tain the non-oxidative use of glucose. We also treated
MCT4-depleted MDA-MB-468 cells with the complex
We next examined whether loss of viability observed I inhibitor metformin. Interestingly, MCT4-depleted
after MCT4 silencing relates to changes in cellular pH cells showed increased sensitivity towards this drug
homeostasis. We assayed the intracellular pH [pH(i)] (Figure 3F; see also supplementary material, Figure
of viable cells after silencing of MCT4. HCC1954 and S3B), indicating that depletion of MCT4 increases the
MDA-MB-468 cells showed a significant reduction dependence of cancer cells on respiration.
in pH(i) upon MCT4 depletion, while no change was As expected, silencing of MCT4 caused a reduction in
observed in MCF10A cells (Figure 3A). Intracellu- lactate secretion (Figure 3G, H; see also supplementary
lar acidification can cause the formation of reactive material, Table S3). Instead, we noticed an increased
oxygen species (ROS), potentially by increasing the uptake of several amino acids, particularly glutamine,
NADH:NAD+ ratio and inhibiting complex I of the res- accompanied by the secretion of glutamate and trypto-
piratory chain [22,23]. HCC1954 and MDA-MB-468 phan (Figure 3G, H). This implies that cells increase
cells showed increased ROS levels after MCT4 ablation. catabolic use of some amino acids, particularly glu-
This was not observed in MCF10A cells (Figure 3B), tamine, when MCT4 function is impaired. This was also
suggesting that these cells do not depend on MCT4 to confirmed by increased sensitivity of MCT4-depleted
maintain pH(i) and prevent ROS formation. cells towards BPTES, a selective inhibitor of glutami-
Lactate production and secretion is important to main- nase, which is essential for the utilization of glutamine
tain non-oxidative glucose metabolism in cancer cells by the TCA cycle (Figure 3I; see also supplementary
[24]. We therefore assayed the effect of MCT1 or material, Figure S3C).
MCT4 ablation on glycolytic flux in MDA-MB-468
cancer cells. Depletion of MCT4 resulted in a 50% MCT4 supports spheroid formation, 3D growth
reduction in glycolytic flux, while depletion of MCT1 and tumour formation in breast cancer cells
had a smaller effect (Figure 3C). We also used stable The regulation of pH(i) by carbonic anhydrase IX (CA9)
isotope labelling to follow changes in metabolite dis- can promote the ability of tumour cells to grow as
tribution. Depletion of MCT4 increased the levels of three-dimensional (3D) spheroids [26]. These culture
M + 2 citrate and M + 2 glutamate in cells labelled with conditions recapitulate the oxygen, nutrient and pH
U-13 C-glucose (Figure 3D), confirming the enhanced gradients experienced by cancer cells within tumours
entry of glucose-derived pyruvate into the TCA cycle. [27]. Analysis of histological sections of spheroids from
In contrast, levels of M + 3 aspartate, M + 5 citrate MDA-MB-468 cells showed that MCT4 is expressed
and M + 4 glutamate were decreased (see supplemen- mainly in a ring surrounding the necrotic core, which
tary material, Figure S3A), suggesting that the direct also stained positive for cleaved caspase 3 (Figure 4A).
conversion of pyruvate into oxaloacetate by pyruvate Silencing of MCT4 caused a significant reduction in
carboxylase (PC) was reduced. We next determined spheroid size in MDA-MB-468 and HCC1954 cells
glucose-dependent respiration using the Seahorse Bio- (Figure 4B), indicating that MCT4 is required for effi-
analyser. We observed a reduction in the oxygen con- cient cell growth under these conditions.
sumption rate (OCR) of MDA-MB-468 cells upon addi- Silencing of MCT4 significantly reduced the
tion of glucose (Figure 3E), suggesting that cells switch growth of MDA-MB-468 and HCC1954 cells in a
from oxidative to non-oxidative glucose metabolism as 3D Matrigel-based model (Figure 4C). In contrast,
glucose becomes available, a phenotype known as the acinar formation of MCF10A cells induced during 14
’Crabtree-effect’ [25]. Silencing of MCT4 reduced the days of 3D culture was insensitive to shRNA-mediated
ability of cancer cells to lower their OCR in response depletion of MCT4 (Figure 4D). Breast epithelial

Figure 1. Identification of metabolic enzymes required for breast cancer cell survival: 17 breast epithelial cell lines were reverse-transfected
with Dharmacon SMARTpool siRNAs targeting 231 different metabolic genes; the cells were placed under normoxic (20% O2 ) or hypoxic
(0.5% O2 ) conditions for 96 h; cell number was used to calculate z-scores for each gene in the different cell lines/conditions. (A) Pie chart
of the metabolic processes represented by the genes targeted by the siGenome siRNAs used in the screen. (B) Supervised cluster analysis of
z-scores derived from cells grown in normoxia; non-malign., non-malignant; genes that showed significant differences in z-score between
the two groups are displayed; p ≤ 0.05. (C) Supervised cluster analysis of z-scores derived from cells grown in hypoxia. (D) Comparison
of z-scores for SLC16A3 (MCT4) between breast cancer subtypes in normoxia and hypoxia. (E) SLC16A3 (MCT4) mRNA expression in
cells grown under normoxic and hypoxic conditions for 24 h; mean ± SD of three biologically independent replicates. (F) Effect of SLC16A3
(MCT4) silencing on cell number in HCC1954, MDA-MB-468 or MCF10A cells under normoxic (black bars) and hypoxic (red bars) conditions;
mean ± SD of two independent experiments with three replicates each; subtypes were compared using Mann–Whitney U-test, *p ≤ 0.05,
**p ≤ 0.01, ***p ≤ 0.001, ****p ≤ 0.0001; n.s., not significant. (G) Differential expression of SLC16A3 (MCT4) in three breast cancer datasets:
significant comparisons between cancerous and normal tissue are shown; p ≤ 0.001; data are publicly available in Oncomine and references
are provided in Supplementary materials and methods (see supplementary material). (H) Differential expression of SLC16A3 (MCT4) in a
study comparing breast cancers that show no recurrence to those that show recurrence at 3 years: statistical comparisons were performed
using Student’s t-test, **p ≤ 0.01, ***p ≤ 0.001, ****p ≤ 0.0001; n.s., not significant

Copyright © 2015 Pathological Society of Great Britain and Ireland. J Pathol 2015; 237: 152–165
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com
MCT4 supports breast cancer cell survival 157

A
SLC16A3 SLC16A1 BSG
15
(normalized to population mean)

(normalized to population mean)


3

(normalized to population mean)


ER+/HER2- ER+/HER2- ER+/HER2-
HER2+ HER2+ HER2+
SLC16A3 mRNA levels

SLC16A1 mRNA levels


ER-/HER2- ER-/HER2- ER-/HER2-
non-malign.

BSG mRNA levels


6 non-malign. non-malign.
10 2

5 1
2

0 0 0

184B5

184B5
BT-549

BT-549
HCC1954

HCC1806

HCC1954

HCC1806
AU-565

AU-565
HS-578T

HS-578T
BT-20

BT-20
184B5

SK-BR-3

SK-BR-3
MCF10A
MCF12A

MCF10A
MCF12A
MDA-MB-468

MDA-MB-231
MDA-MB-436

MDA-MB-468

MDA-MB-231
MDA-MB-436
ZR-75-1

ZR-75-1
T-47D

HCC38

T-47D

HCC38
BT-549
HCC1954

HCC1806

MCF7

MCF7
AU-565

HS-578T
BT-20
SK-BR-3

MCF10A
MCF12A
MDA-MB-468

MDA-MB-231
MDA-MB-436
ZR-75-1

T-47D

HCC38
MCF7

-5 B- 1
6
B C

68

BT M 3
H 49 43
****

A- -2
-2 B-4

C 6
54
n.s.
(normalized to population mean)

D B
D T
- 3
(normalized to population mean)

C A
S A

C 0
15

M 5-1

M 578
M -M
H 565
AU R-

H 18
BT -M

R 38
R C19
10

M 10
R 2
****

M 5
SK D
T- F7

18 0

F1
-B

4B

A
-7

C
A

F
47

S-
****
SLC16A1 mRNA levels

C
C
SLC16A3 mRNA levels

ZR

S
S
M

H
****
8
*** 10 MCT4
**
6
MCT1
4
5

2
BSG
0
β-actin
0
ER .
H n.

+/ lign

ER+/HER2-
H 2-
-/H 2+
2-
ER ER -
-/H 2+

ER-/HER2-
2-
H 2
ig

ER
ER

ER

ER ER
al

HER2+ non-malign.
ER -ma
m

H
n-

+/

n
no
no
ER

Figure 2. Expression of SLC16A3 and SLC16A1 in breast cancer cells. (A) mRNA levels of SLC16A3 (MCT4), SLC16A1 (MCT1) and BSG in
breast epithelial cell line panel. (B) Comparison of mRNA levels of SLC16A3 (MCT4) and SLC16A1 (MCT1) between different breast cancer
subtypes; non-malign., non-malignant. (C) Protein levels of MCT4, MCT1 and BSG across breast epithelial cell line panel: β-actin is shown
as loading control; a reference sample (RS) was loaded to allow comparison of signal intensities across multiple gels; Student’s t-test,
**p ≤ 0.01, ***p ≤ 0.001, ****p ≤ 0.0001; n.s., not significant

cells alter their metabolic activity in response to loss material, Figure S1A). Moreover, analysis of the TCGA
of extracellular matrix (ECM) attachment [28] and dataset showed that SLC16A3 (MCT4) amplifications,
MCT4 seems to be required for this adaptation. We next which occur in 6.4% of breast cancers, have a tendency
investigated whether co-culture with cancer-associated towards co-occurrence with HER2/ERBB2 amplifica-
fibroblasts (CAFs) alters the sensitivity of breast can- tion (p = 0.000292) (MSKCC cBioPortal). Most breast
cer cells towards ablation of MCT4. Transfection of cancers harbouring these genetic alterations display
species-specific siRNA targeting human MCT4 caused high levels of Akt phosphorylation [10].
a substantial loss of viability (Figure 4E) and induc- To investigate whether MCT4 dependence is asso-
tion of apoptosis (Figure 4F) in MDA-MB-468 cells ciated with activation of the PI3K–Akt pathway, we
cultured either alone or in the presence of CAFs. This established levels of growth factor-independent Akt
demonstrates that MCT4 has essential functions in phosphorylation in all cell lines. High levels of Akt ser-
cancer cells, even under conditions that allow the shut- ine 473 phosphorylation were detected in ER− HER2−
tling of metabolites between cancer and stromal cells. breast cancer cells (see supplementary material, Figure
Moreover, inducible silencing of MCT4 substantially S4) and most cancer cells maintained high Akt phospho-
reduced the growth of MDA-MB-468 cells as orthotopic rylation in the absence of growth factors (Figure 5A;
tumours (Figure 4G, H), confirming the importance of see also supplementary material, Figure S4). We next
MCT4 for breast cancer cells under physiological compared levels of growth factor-independent Akt phos-
growth conditions. phorylation across the breast cancer cell line panel with
their sensitivity towards MCT4 ablation. Cell lines that
MCT4 is regulated by the PI3K–Akt signalling axis retained high levels of Akt phosphorylation, such as
in breast cancer MDA-MB-468 or HCC1954, displayed high sensitivity
Having established that MCT4 is essential for the towards MCT4 silencing (ie low z-scores; Figure 5B),
growth and survival of breast cancer cell lines, we suggesting a functional link between MCT4 and the
investigated whether it is linked to specific genetic PI3K–Akt signalling axis.
alterations found in the human disease. Sensitivity We next investigated whether activation of the
towards MCT4 depletion was most pronounced in PI3K–Akt signalling pathway regulates MCT4 expres-
the HER2+ subset or those cell lines that harbour sion. We chose two breast cancer cell lines that show
mutations in PTEN or INPP4B (see supplementary amplifications within EGFR (MDA-MB-468) or HER2
Copyright © 2015 Pathological Society of Great Britain and Ireland. J Pathol 2015; 237: 152–165
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com
158 F Baenke et al

A B
HCC1954 MDA-MB-468 MCF10A HCC1954 MDA-MB-468 MCF10A
** 2.5
8.5 8.5 2.5 6
8.5

H2DCFDA fluorescence
H2DCFDA fluorescence

H2DCFDA fluorescence
2.0

(relative to siCtrl)
**** ** 2.0

(relative to siCtrl)

(relative to siCtrl)
8.0 n.s.
8.0 8.0 n.s.
4 1.5
1.5 ****

pH(i)
pH(i)

pH(i)

7.5 7.5 7.5


1.0 1.0
2
7.0 7.0 7.0 0.5
0.5

6.5 0.0 0 0.0


6.5 6.5

T4
2
+ 4
si trl
2
T4
s trl

si rl
2O

2
2O

T
T4
T4

t
T4

trl

2O
C

C
trl
trl

C
C
si iMC

H
si

M
C

H
C

si

si
C

M
C
C

H
si
si
si

M
M
M

+
trl
si
si
si

trl

trl
C
C

C
si

si
C D U-13C-Glucose Citrate (M+2) Glutamate (M+2)
MDA-MB-468 - Dox 25 p=0.0052
- Dox
1.5 p=0.00027
+ Dox
+ Dox
relative 3H2O production

% of total pool
*** 30

% of total pool
Pyruvate 20
(relative to siCtrl)

**
15
1.0 20
Acetyl-CoA
10
Citrate 10
Oxaloacetate 5
0.5
0 0

8
R
8
R

#8
Glutamate

#8

SC
SC

0.0 Malate α-Ketoglutarate

T4
T4

sh
sh

C
T1

T4
trl

M
C

M
C

C
si

sh
M

sh
si

si

E shMCT4 #74 shMCT4 #88 F shMCT4 #88


shSCR
110 glucose 110 glucose 110 glucose
***
OCR [pM/min]/cell mass
OCR [pM/min]/cell mass

OCR [pM/min]/cell mass

- Dox - Dox - Dox


absorbance (560 nm) - Dox
100 100 100 (normalized to -Dox) 1.0 + Dox
+ Dox + Dox + Dox
90 90 90
80 ***
80 **** 80
***
70 70 70 0.5
** ***
60 60 60
***
50 50 50
40 40 0.0
40
8 14 21 28 35 43
in

8 14 21 28 35 43
in
8 14 21 28 35 43
t

rm
en

rm

time [min] time [min] time [min]


fo
lv

fo
so

et

et
m

G H
M

M
1m

5m

Lactate Glutamine
15 2
Lactate siCtrl
nM/(cell mass unit × h)

nM/(cell mass unit × h)

Malate siMCT4
1
Fumarate
10
Succinate
α-Ketoglutarate
0
I shMCT4 #88
Citrate -1 ***
5 - Dox
absorbance (560 nm)
(normalized to -Dox)

Alanine 1.0 + Dox


Asparagine -2
***
Aspartate ***
0 -3
Glutamate ***
trl
T4

T4
M rl
t
C

Glutamine 0.5
C
C

C
si

si
M

Glycine
si

si

Isoleucine
Glutamate Tryptophan
Leucine
0.25 1.5
Methionine 0.0
nM/(cell mass unit × h)

nM/(cell mass unit × h)

Phenylalanine
t

S
en

0.20
TE

TE
lv

Proline
BP

BP
so

1.0
Serine 0.15
M

μM

Succinate
10

Tryptophan 0.10
0.5
Tyrosine
0.05
Valine
-1.0 -0.5 0.0 0.5 1.0 10 15 20 0.00 0.0
-3 5
-2 0
-2 5
-1 0
.5
.
.
.
.
-3

trl
T4

nM/(cell mass unit × h)


trl
T4

C
C

C
si
C
si

M
M

si
si

Figure 3. Legend on the next page

Copyright © 2015 Pathological Society of Great Britain and Ireland. J Pathol 2015; 237: 152–165
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com
MCT4 supports breast cancer cell survival 159

(HCC1954) and have high levels of Akt activation Discussion


(Figure 5A). EGF increased MCT4 mRNA levels in
both cell lines (Figure 5C). Moreover, treatment with the Metabolic reprogramming is a hallmark of cancer and
HER2/EGFR inhibitor lapatinib or inhibitors of PI3K may offer novel treatment opportunities [4]. Here we
(PI-103 or LY294002) attenuated MCT4 induction, demonstrate the role of the monocarboxylate transporter
while inhibitors of MEK–ERK signalling (PD98059 MCT4 in supporting cancer cell survival in a panel of
or UO126) were less effective (Figure 5C). Finally, genetically diverse breast cancer cell lines. We found
induction of MCT4 expression by hypoxia in MCF10A that MCT4 is required for metabolite transport, pH
cells was dependent on the presence of growth factors, homeostasis and redox balance, and that it supports the
including insulin and EGF (Figure 5D). viability of breast cancer cells under physiologically rel-
evant growth conditions in vitro and in a breast cancer
MCT4 expression is induced in breast cancer xenograft model. Moreover, expression of MCT4 is reg-
The MMTC–PyMT mouse model of breast cancer reca- ulated by the PI3K–Akt signalling axis and correlates
pitulates the different stages of breast cancer progres- with disease progression in a mouse model of breast can-
sion [30] and is highly dependent on Akt activity cer. High levels of MCT4 are found in human breast
[31]. We therefore investigated expression of MCT4 cancer tissue and are predictive of poor survival.
in tissues from these mice at different disease stages. There is substantial evidence that proton-linked
MCT4 expression was low in normal murine mam- monocarboxylate transporters perform important roles
mary glands but increased with neoplastic transfor- in cancer cells [32]. The low affinity of MCT4 to pyru-
mation (Figure 6A). In the carcinoma stage, strong vate facilitates the conversion of pyruvate into lactate
staining was localized to cancer cells in areas surround- in hypoxic cancer cells, thereby allowing the regener-
ing necrotic tissue, suggesting that MCT4 is induced by ation of cytosolic NAD+ [33]. In contrast, MCT1 has
tumour hypoxia and/or oxidative stress. high affinity for both lactate and pyruvate and may be
We next analysed MCT4 expression in a tissue more important for lactate uptake in oxidative cancer
microarray of 75 human breast specimens. MCT4 cells. However, it is now clear that the expression of
staining was low in normal breast (Figure 6B) but MCT1 and MCT4 can differ substantially, not only
enhanced in cancerous tissues from all clinical subtypes between different cancer types but also between dif-
(Figure 6C, E). Moreover, some HER2+ and ER+ breast ferent intra-tumoural compartments and cell types.
cancers showed high-intensity MCT4 staining localized Functional studies are therefore needed to unravel the
to the cancer compartment (Figure 6C, E). In contrast, relative contributions of different lactate transporters.
stromal expression of MCT4 was moderate and uniform It has been shown that combined inhibition of the
across all subtypes (Figure 6D, F). A meta-analysis of lactate transporters MCT4 and MCT1 blocks viability
expression data showed that high levels of SCL16A3 and tumour growth in RAS-transformed fibroblasts and
are indicative of poor overall survival in HER2+ breast LS174T colon cancer cells [19]. However, we found
cancer (Figure 6G). Collectively, these results indicate that ablation of MCT4 alone was sufficient to impair
that MCT4 is regulated by the PI3K–Akt signalling viability of human breast cancer cell lines. Sensitivity
axis in human breast cancer and promotes the survival towards MCT4 ablation did not correlate with expres-
of breast cancer cells in which this pathway is activated. sion of either MCT4 or MCT1, making it unlikely

Figure 3. MCT4 is required for pH maintenance, redox balance and non-oxidative glucose metabolism in breast cancer cells. (A) HCC1954,
MDA-MB-468 and MCF10A cells were transfected with siRNA targeting MCT4 (siMCT4) or controls (siCtrl): 72 h post-transfection,
intracellular pH [pH(i)] was analysed; mean ± SD of two independent experiments with two replicates each. (B) ROS levels in cells treated as
in (A): H2 O2 treatment was used as positive control for ROS induction; mean ± SEM of three independent experiments with two replicates
each. (C) Effect of MCT1 or MCT4 ablation on glycolytic activity in MDA-MB-468 cells; mean ± SEM of two independent experiments
with four replicates each. (D) MDA-MB-468 cells expressing inducible shRNAs targeting MCT4 (shMCT4 88) or scrambled controls (shSCR)
were grown in the presence or absence of doxycycline (± Dox) for 7 days: cells were labelled with U-13 C-glucose for 24 h before lysis and
metabolite levels were determined by mass spectrometry; increased levels of labelled citrate (M + 2) and glutamate (M + 2) are indicative
of enhanced entry of pyruvate into the TCA cycle. (E) MDA-MB-468 cells expressing inducible shRNAs targeting MCT4 (shMCT4 74 or
shMCT4 88) or scrambled controls (shSCR) were grown in the presence or absence of doxycycline (± Dox) for 6 days: the cells were starved
of glucose and glutamine for 2 h and oxygen consumption rate (OCR) was assessed before and after injection of 10 mM glucose, using a
Seahorse Bioanalyser; mean ± SD of two independent experiments with six replicates each. (F) MDA-MB-468 cells expressing inducible
shRNAs targeting MCT4 (shMCT4 88) were grown in the presence or absence of doxycycline (± Dox) for 8 days: the indicated doses of
metformin were added during the last 3 days; cell viability was determined by crystal violet staining; mean ± SEM of two independent
experiments with three replicates each. (G) Effect of MCT4 ablation on metabolite uptake and secretion in MDA-MB-468 cells under
normoxia: cells were transfected with siRNA targeting MCT4 (siMCT4) or controls (siCtrl); 72 h post-transfection, cells were placed in fresh
medium and samples of the medium were taken over a period of 24 h; metabolite levels in the medium were determined using GC–MS;
data represent six biological replicates. (H) Uptake and secretion rates of lactate, glutamine, glutamate and tryptophan in MDA-MB-468
cells, as in (G). (I) MDA-MB-468 cells expressing inducible shRNAs targeting MCT4 (shMCT4 88) were grown in the presence or absence of
doxycycline (± Dox) for 8 days: the indicated doses of BPTES were added during the last 3 days; cell viability was determined by crystal violet
staining; mean ± SEM of two independent experiments with three replicates each; Student’s t-test, **p ≤ 0.01, ***p ≤ 0.001, ****p ≤ 0.0001;
n.s., not significant

Copyright © 2015 Pathological Society of Great Britain and Ireland. J Pathol 2015; 237: 152–165
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com
160 F Baenke et al

A B MDA-MB-468 HCC1954
MCT4 cleaved caspase 3
siCtrl 2000 siCtrl 1500

spheroid area [pixelx1000]


spheroid area [pixelx1000]
**** ****
1500
1000

100 μm
1000 100 μm
100 μm 100 μm
siMCT4 siMCT4
500
500

0 0

trl

T4
100 μm 100 μm

T4
trl

C
C

si

M
si

si
si
100 μm

H&E
C D
MDA-MB-468 HCC1954
siCtrl siCtrl MCF10A shSCR

1.5 1.5 1.5

(normalized to siCtrl)
(normalized to siCtrl)

(normalized to siCtrl)
****
****

cell viability
cell viability

1.0
cell viability

1.0 1.0 + Dox


siMCT4 siMCT4 shMCT4 #88

0.5 0.5 0.5

0.0 0.0 0.0 +Dox


siCtrl siMCT4 siCtrl siMCT4 shSCR shMCT4
(#88)

MDA-MB-468 siCtrl MDA-MB-468 + CAFs siCtrl


E MDA-MB-468 MDA-MB-468 + CAFs F 5 0.39% 1.68% 0.86% 3.12%
10 105
0.39%
4
10 104

PI-A
PI-A

siCtrl 10 3 103

10 2 102

0 0
94.1% 3.38% 90.5% 5.51%
0 10
2
10
3 4
10 10
5 0 102 103 104 105
Annexin V Pacific Blue Annexin V Pacific Blue
MDA-MB-468 siMCT4 MDA-MB-468 + CAFs siMCT4
0.7% 1.31% 105
0.96% 1.58%
105
siMCT4

4
104 10
PI-A

PI-A

3
103 10

G MDA-MB-468 shMCT4 #88 102 102


150 - Dox 0
0
+ Dox 74.6% 23.4% 75.6% 21.8%
tumour volume (mm )
3

2 3 4 5 2 3 4 5
0 10 10 10 10 0 10 10 10 10
Annexin V Pacific Blue Annexin V Pacific Blue
100
H - Dox + Dox

50 Dox added *

0
0 20 40 60 days

Figure 4. MCT4 depletion reduces 3D growth and tumour formation in breast cancer cells. (A) MDA-MB-468 cells were grown as spheroids
for 7 days; spheroids were sectioned and stained for MCT4 or cleaved caspase 3; H&E staining is shown as control. (B) MDA-MB-468 and
HCC1954 cells were transfected with siRNA targeting MCT4 (siMCT4) or controls (siCtrl): 12 h post-transfection, cells were trypsinized,
seeded as spheroids and incubated for 7 days; spheroid size was determined and compared to siCtrl-transfected cells; Student’s t-test,
****p ≤ 0.0001. (C) MDA-MB-468 and HCC1954 cells were transfected with siRNA targeting MCT4 (siMCT4) or controls (siCtrl): 12 h
post-transfection, cells were trypsinized and embedded in Matrigel; after 7 days, representative images were taken and viable cell number
determined with CellTiter-Blue, normalized to siCtrl-transfected cells; mean ± SEM of three independent experiments with two replicates
each; Student’s t-test, ****p ≤ 0.0001. (D) MCF10A cells expressing shMCT4 88 or scrambled controls (shSCR) were embedded in Matrigel
and treated with 1 μg/ml doxycycline for 14 days: representative images were taken and viable cell number determined with CellTiter-Blue;
mean ± SEM of three independent replicates. (E) Representative images of co-cultures of MDA-MB-468 with mCherry-labelled murine
cancer-associated fibroblasts (CAFs), 3 days after transfecting with siRNA targeting MCT4 (siMCT4) or control (siCtrl); arrows, apoptotic
MDA-MB-468 cells. (F) Apoptosis in the mCherry-negative populations (= MDA-MB-468 cells), determined by Annexin V–PI staining and
FACS. (G) MDA-MB-468 cells expressing shMCT4 88 were injected into the mammary fat pad of 8 week-old immunocompromised mice
(nu/nu): after tumour initiation on day 10, the mice were randomized into two cohorts (n = 5); one cohort was kept on a standard diet
and the other on a doxycycline-supplemented diet (0.2 g/kg doxycycline in food pellet); tumour size was measured twice weekly; statistical
comparisons of sizes at day 65 were performed using Student’s t-test, *p ≤ 0.05. (H) Representative images of tumour tissue

Copyright © 2015 Pathological Society of Great Britain and Ireland. J Pathol 2015; 237: 152–165
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com
MCT4 supports breast cancer cell survival 161

A 5% HS + GF
2.5 5% HS - GF
phospho-AKT/pan-AKT levels

10% FCS
(normalized to β-actin)

0.5% BSA
2.0

1.5

1.0

0.5

non-malign.
0.0 ER+/HER2-
HER2+

38
8
5

M 1

BT 6
AU -3
M 0A

ZR 2A

M CC 5
M 4

H 6
C 49
H -20

M MB T
F7

SK 7D
1

ER-/HER2-
46
4B

A- -23

43
6

0
5-

C
R

5
D 19

18
5
D -57
C
F1

F1

C
B-

BT
-7

-B

B-
-

-
18

T-
M

C
C

S
M

A-

A-
H

H
D
M
B 2 r = -0.598 C
184B5 p = 0.012 HCC1954 MDA-MB-468
M12A 2.5 ** 5
M-231
SLC16A3 mRNA levels

SLC16A3 mRNA levels


(normalized to control)

(normalized to control)
0 BT-20 ****
ZR-75-1
SLC16A3 z-score

M10A HS-578T 2.0


**
4
MCF7 **
M-436 H38
-2
AU-565 T-47D 1.5 *** ** 3 **
BT-549 M-468 ***
SK-BR-3 H1806 *** ***
1.0 2
-4 H1954
0.5 1

-6 0.0 0
0.0 0.5 1.0 1.5 2.0 2.5
F a

EG 103

LY

F a

EG 103

LY
EG F D

EG F D
l

l
+ F

+ F
O

O
ro

ro
EG ap

EG ap
EG

EG
EG + P

EG + P
F +U

F +U
nt

nt
+

+
L

-
pAKT in 0.5% BSA
co

co
PI

PI
F

F
+

+
F

F
(normalized to β-actin)
EG

EG
D MCF10A
2.5 - GF
(normalized to 20% O2 + GF)

**
+ GF
SLC16A3 mRNA levels

2.0

1.5
**
1.0

0.5

0.0
20% O2 0.5% O2

Figure 5. MCT4 expression is regulated by EGFR–HER2 signalling and increases during breast cancer progression. (A) Cells were cultured
in medium containing 10% FCS or 0.5% bovine serum albumin (BSA); non-malignant cells were cultured in medium with 5% horse serum
(HS), with or without growth factor supplements (GF): Akt phosphorylation levels were determined by immunoblotting (see supplementary
material, Figure S4); signal intensities were analysed using ImageJ to determine the phospho-AKT:pan-AKT ratio. (B) Akt phosphorylation
levels in medium with 0.5% BSA were compared to z-scores determined after MCT4 silencing, using the Spearman correlation. (C) HCC1954
and MDA-MB-468 cells were serum-starved for 16 h and stimulated with 20 ng/ml EGF for 6 h, with or without 1 h pretreatment with the
HER2/EGFR inhibitor lapatinib (Lapa; 0.5 μM), the MEK inhibitors PD98059 (PD; 10 μM) or UO126 (UO; 10 μM), the dual PI3K–mTOR inhibitor
PI-103 (1 μM) or the PI3K inhibitor LY294002 (LY; 20 μM). SLC16A3 (MCT4) mRNA levels were determined by qPCR; mean ± SEM of two
independent experiments with two replicates each. (D) MCF10A cells cultured in medium with (+GF) or without (–GF) growth factors
under normoxia (20% O2 ) or hypoxia (0.5% O2 ) for 24 h: expression of SLC16A3 (MCT4) mRNA was determined by qPCR; mean ± SD of two
replicates

that MCT1 compensates for MCT4 in breast cancer. We found that reduced viability after MCT4 deple-
Another study reported that 7-aminocarboxycoumarin tion was associated with intracellular acidification and
derivatives, which inhibit lactate uptake but not secre- oxidative stress, indicating that MCT4 is required for
tion, block xenograft growth of several cancer cell lines, the maintenance of pH homeostasis and redox bal-
including ER+ MCF7 cells [34]. Here ER+ breast cancer ance. MCT4 depletion reduced glycolytic flux in cancer
cells showed low sensitivity towards MCT4 depletion cells but increased the entry of glucose-derived pyru-
compared to other subtypes, suggesting that lactate vate into the TCA cycle. Consistent with the function
metabolism differs substantially between breast cancer of MCT4 as a lactate–proton symporter, we observed
cell lines. a reduction in lactate secretion after MCT4 depletion.
Copyright © 2015 Pathological Society of Great Britain and Ireland. J Pathol 2015; 237: 152–165
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com
162 F Baenke et al

adenoma
A normal hyperplasia early carcinoma carcinoma

100μm 100μm 100μm 100μm

B normal mammary gland C cancer D stromal MCT4


intensity 0 intensity 1 intensity 1

100μm 100μm 100μm 100μm

intensity 2 intensity 3 intensity 2

100μm 100μm 100μm

E F G
cancer HER2-positive breast cancer
staining stroma staining 100
100 intensity intensity
100
0 1 80
80
overall survival (%)

1 80 2
p= 0.0314
percentage

2
percentage

60
60 60
3
40
40 p = 0.03 40
20 low MCT4
20 20
med MCT4
0 high MCT4
0 0
0 2 4 6 8 10
)
)

)
)
31
30

35
24

)
1)

24

35
30

Time (years)
=
=

=
=

=
=
=
(n
(n

(n
(n

(n

(n
(n
(n
+

3+
BC

2+

2+

3+
BC
ER

ER
2
TN

2
TN
ER
ER

ER

ER
H
H

Figure 6. MCT4 is expressed in human breast cancer and predicts poor survival in HER2-positive disease. (A) Breast tissues derived from
MMTV–PyMT mice at different stages of tumour development were stained for MCT4: expression of MCT4 in breast tissue was analysed
using a tissue microarray; representative images are shown. (B) Normal mammary gland. (C) Different staining intensities used for scoring
cancer cell positivity. (D) Different staining intensities used for scoring stromal cell positivity. (E) Percentage of tissue cores with different
staining intensities for MCT4 in cancer cells within the different subgroups, according to receptor status: staining intensities for HER2 and
ER according to the manufacturer’s information; HER2 2+ = equivocal, HER2 3+ = HER2 amplified; p value reflects c2 statistical significance
in the correlation analysis. (F) Percentage of tissue cores with different staining intensities for MCT4 in stromal cells, according to receptor
status. (G) Kaplan–Meier curves showing the overall survival of patients with HER2-positive breast cancers, showing low (grey line), medium
(black line) or high (red line) levels of SLC16A3 (MCT4) mRNA expression

Instead, cells increased the uptake of amino acids, partic- and glutaminolysis, indicating that cells adapt to MCT4
ularly glutamine, suggesting an induction of anaplerosis inhibition by increasing oxidative metabolism. Cancer
and use of amino acids for energy generation. Inter- cells that are unable to convert to oxidative metabolism
estingly, MCT4-depleted cells showed enhanced sen- should therefore be particularly sensitive to MCT4 inhi-
sitivity towards inhibitors of mitochondrial respiration bition. Similar observations were also recently reported
Copyright © 2015 Pathological Society of Great Britain and Ireland. J Pathol 2015; 237: 152–165
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com
MCT4 supports breast cancer cell survival 163

for a glycolytic subtype of pancreatic cancer [35], sug- We found that a subset of breast cancers showed mod-
gesting that combination strategies targeting glutaminol- erate stromal MCT4 staining, independent of receptor
ysis or mitochondrial respiration, together with MCT4 status. In contrast, strong positivity for MCT4 in can-
inhibitors, could be of clinical benefit. cer cells was found in a proportion of tumours, with
As expected, sensitivity of breast cancer cells towards the strongest staining detected in HER2+ and ER+ sub-
MCT4 depletion was more pronounced when cells were types. It is likely that the regulation of MCT4 expression
grown in hypoxia. MCT4 is induced by hypoxia in a in breast cancer is highly dynamic. Cancer cells need
HIF-1a-dependent manner [29]. It is likely that ablation to respond to different metabolic niches determined by
of MCT4 lowers the ability of breast cancer cells to adapt the tumour microenvironment. Differential expression
their metabolism to hypoxia. Consequently, MCT4 was of MCTs allows cancer cells to modify their metabolic
also required to support the growth of breast cancer cells activity according to the nutrient and oxygen conditions
as 3D spheroids, which recapitulate oxygen, nutrient and and their metabolic demand. Understanding the depen-
pH gradients experienced by cancer cells within growing dencies of cancer cells within the context of this dynamic
tumours [27]. regulation is essential to develop strategies to target lac-
Several studies have shown that metabolic reprogram- tate transport for the treatment of cancer.
ming contributes to the morphological transformation
of breast cancer cells [28,36,37]. Cancer phenotypes
such as loss of cell polarization, defective morphogen-
Acknowledgements
esis and enhanced invasion can be investigated in 3D
collagen cultures, and alter the sensitivity of cancer
cells towards inhibition of signalling molecules [38]. We The authors thank Eric Sahai (LRI) for material, Tamara
found that depletion of MCT4 reduced the 3D growth Bunting for tissue analysis and LRI Research Services
of breast cancer cells but not non-malignant cells, sug- for technical support. This study was funded by Cancer
gesting that MCT4 supports the metabolic adaptation of Research UK and the Comprehensive Cancer Centre
cancer cells to these conditions. It has been suggested Mainfranken.
that MCT4 participates in the shuttling of lactate from
glycolytic stromal cells to support oxidative metabolism Author contributions
in cancer cells [39]. We found that co-culture with
cancer-associated fibroblasts (CAFs) did not alter the FB, SD, CB, BD, BG, MJ and SR designed and carried
sensitivity of cancer cells towards MCT4 depletion. out experiments; BW, JRF, MH, NZ and AS conceived
Instead, we found that MCT4 is essential for orthotopic the study; FB and AS wrote the manuscript; AM and
tumour formation, confirming that MCT4 is required for BS performed statistical analysis; and BSD and GS
cancer cell growth under physiological growth condi- performed histological analysis.
tions.
Our study also revealed a regulatory connection
between MCT4 and PI3K–Akt signalling in breast
cancer. Signalling through the PI3K–Akt axis also References
results in enhanced HIF-1 activity in normoxic cancer 1. Vander Heiden MG, Cantley LC, Thompson CB. Understanding
cells [41]. It is therefore possible that the induction of the Warburg effect: the metabolic requirements of cell proliferation.
MCT4 expression in response to EGF involves activa- Science 2009; 324: 1029–1033.
2. Lunt SY, Vander Heiden MG. Aerobic glycolysis: meeting the
tion of HIF-1. Akt increases glucose uptake and lactate
metabolic requirements of cell proliferation. Annu Rev Cell Dev Biol
secretion in cancer cells [40] and impairment of lactate
2011; 27: 441–464.
transport could be highly detrimental to cells with active 3. Cairns RA, Harris IS, Mak TW. Regulation of cancer cell
Akt. The connection between MCT4 and PI3K–Akt metabolism. Nat Rev Cancer 2011; 11: 85–95.
was confirmed by the observation that MCT4 levels 4. Schulze A, Harris AL. How cancer metabolism is tuned for prolifer-
increase with disease progression in the MMTV–PyMT ation and vulnerable to disruption. Nature 2012; 491: 364–373.
breast cancer model, which shows enhanced HER2 5. Tennant DA, Duran RV, Gottlieb E. Targeting metabolic transforma-
expression and Akt phosphorylation during progression tion for cancer therapy. Nat Rev Cancer 2010; 10: 267–277.
to the carcinoma stage [42]. Consistent with a role 6. Galluzzi L, Kepp O, Vander Heiden MG, et al. Metabolic targets for
of MCT4 in supporting the metabolism of hypoxic cancer therapy. Nat Rev Drug Discov 2013; 12: 829–846.
cancer cells [43], staining was mainly localized to areas 7. Parker JS, Mullins M, Cheang MC, et al. Supervised risk predictor
surrounding necrotic tissue. of breast cancer based on intrinsic subtypes. J Clin Oncol 2009; 27:
1160–1167.
MCT4 staining shows substantial heterogeneity in
8. Reis-Filho JS, Pusztai L. Gene expression profiling in breast cancer:
different tumour compartments, including stromal cells
classification, prognostication, and prediction. Lancet 2011; 378:
[21,39,44–46]. MCT4 expression in cancer cells is 1812–1823.
found in triple-negative breast cancer [47] and corre- 9. Curtis C, Shah SP, Chin SF, et al. The genomic and transcriptomic
lates with high-grade disease, while stromal expression architecture of 2000 breast tumours reveals novel subgroups. Nature
was mainly restricted to lower grades [44]. Another 2012; 486: 346–352.
study found that stromal MCT4 expression predicts poor 10. TCGA. Comprehensive molecular portraits of human breast tumours.
disease outcome in triple-negative breast cancer [48]. Nature 2012; 490: 61–70.

Copyright © 2015 Pathological Society of Great Britain and Ireland. J Pathol 2015; 237: 152–165
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com
164 F Baenke et al

11. Shah SP, Roth A, Goya R, et al. The clonal and mutational evolution 30. Lin EY, Jones JG, Li P, et al. Progression to malignancy in the
spectrum of primary triple-negative breast cancers. Nature 2012; 486: polyoma middle T oncoprotein mouse breast cancer model pro-
395–399. vides a reliable model for human diseases. Am J Pathol 2003; 163:
12. Lehmann BD, Bauer JA, Chen X, et al. Identification of human 2113–2126.
triple-negative breast cancer subtypes and preclinical models for 31. Maroulakou IG, Oemler W, Naber SP, et al. Akt1 ablation
selection of targeted therapies. J Clin Invest 2011; 121: 2750–2767. inhibits, whereas Akt2 ablation accelerates, the development of
13. Kan Z, Jaiswal BS, Stinson J, et al. Diverse somatic mutation pat- mammary adenocarcinomas in mouse mammary tumor virus
terns and pathway alterations in human cancers. Nature 2010; 466: (MMTV)-ErbB2/neu and MMTV-polyoma middle T transgenic
869–873. mice. Cancer Res 2007; 67: 167–177.
14. Stephens PJ, Tarpey PS, Davies H, et al. The landscape of cancer 32. Brahimi-Horn MC, Bellot G, Pouyssegur J. Hypoxia and energetic
genes and mutational processes in breast cancer. Nature 2012; 486: tumour metabolism. Curr Opin Genet Dev 2011; 21: 67–72.
400–404. 33. Dimmer KS, Friedrich B, Lang F, et al. The low-affinity monocar-
15. Ward PS, Thompson CB. Metabolic reprogramming: a cancer hall- boxylate transporter MCT4 is adapted to the export of lactate in
mark even Warburg did not anticipate. Cancer Cell 2012; 21: highly glycolytic cells. Biochem J 2000; 350 Pt 1: 219–227.
297–308. 34. Draoui N, Schicke O, Seront E, et al. Antitumor activity of
16. Axelson H, Fredlund E, Ovenberger M, et al. Hypoxia-induced ded- 7-aminocarboxycoumarin derivatives, a new class of potent
ifferentiation of tumor cells – a mechanism behind heterogeneity inhibitors of lactate influx but not efflux. Mol Cancer Ther 2014; 13:
and aggressiveness of solid tumors. Semin Cell Dev Biol 2005; 16: 1410–1418.
554–563. 35. Baek G, Tse YF, Hu Z, et al. MCT4 defines a glycolytic subtype
of pancreatic cancer with poor prognosis and unique metabolic
17. Ward C, Langdon SP, Mullen P, et al. New strategies for targeting
dependencies. Cell Rep 2014; 9: 2233–2249.
the hypoxic tumour microenvironment in breast cancer. Cancer Treat
36. Grassian AR, Metallo CM, Coloff JL, et al. Erk regulation of pyru-
Rev 2013; 39: 171–179.
vate dehydrogenase flux through PDK4 modulates cell proliferation.
18. Lee GY, Kenny PA, Lee EH, et al. Three-dimensional culture models
Genes Dev 2011; 25: 1716–1733.
of normal and malignant breast epithelial cells. Nat Methods 2007;
37. Locasale JW, Grassian AR, Melman T, et al. Phosphoglycerate dehy-
4: 359–365.
drogenase diverts glycolytic flux and contributes to oncogenesis. Nat
19. Le Floch R, Chiche J, Marchiq I, et al. CD147 subunit of lactate/H+
Genet 2011; 43: 869–874.
symporters MCT1 and hypoxia-inducible MCT4 is critical for ener-
38. Weigelt B, Lo AT, Park CC, et al. HER2 signaling pathway activation
getics and growth of glycolytic tumors. Proc Natl Acad Sci USA
and response of breast cancer cells to HER2-targeting agents is
2011; 108: 16663–16668.
dependent strongly on the 3D microenvironment. Breast Cancer Res
20. Chiche J, Fur YL, Vilmen C, et al. In vivo pH in metabolic-defective
Treat 2010; 122: 35–43.
Ras-transformed fibroblast tumors: key role of the monocarboxylate
39. Whitaker-Menezes D, Martinez-Outschoorn UE, Lin Z, et al. Evi-
transporter, MCT4, for inducing an alkaline intracellular pH. Int J
dence for a stromal-epithelial ’lactate shuttle’ in human tumors:
Cancer 2012; 130: 1511–1520.
MCT4 is a marker of oxidative stress in cancer-associated fibroblasts.
21. Pinheiro C, Albergaria A, Paredes J, et al. Monocarboxylate trans- Cell Cycle 2011; 10: 1772–1783.
porter 1 is up-regulated in basal-like breast carcinoma. Histopathol- 40. Elstrom RL, Bauer DE, Buzzai M, et al. Akt stimulates aerobic
ogy 2010; 56: 860–867. glycolysis in cancer cells. Cancer Res 2004; 64: 3892–3899.
22. Riemann A, Schneider B, Ihling A, et al. Acidic environment leads 41. Semenza GL. HIF-1 mediates metabolic responses to intratu-
to ROS-induced MAPK signaling in cancer cells. PloS One 2011; 6: moral hypoxia and oncogenic mutations. J Clin Invest 2013; 123:
e22445. 3664–3671.
23. Lee I, Lee SJ, Kang WK, et al. Inhibition of monocarboxylate trans- 42. Wang S, Yuan Y, Liao L, et al. Disruption of the SRC-1 gene in mice
porter 2 induces senescence-associated mitochondrial dysfunction suppresses breast cancer metastasis without affecting primary tumor
and suppresses progression of colorectal malignancies in vivo. Mol formation. Proc Natl Acad Sci USA 2009; 106: 151–156.
Cancer Ther 2012; 11: 2342–2351. 43. Sonveaux P, Vegran F, Schroeder T, et al. Targeting lactate-fueled
24. Fantin VR, St-Pierre J, Leder P. Attenuation of LDH-A expression respiration selectively kills hypoxic tumor cells in mice. J Clin Invest
uncovers a link between glycolysis, mitochondrial physiology, and 2008; 118: 3930–3942.
tumor maintenance. Cancer Cell 2006; 9: 425–434. 44. Choi J, Kim do H, Jung WH, et al. Metabolic interaction between
25. Diaz-Ruiz R, Rigoulet M, Devin A. The Warburg and Crabtree cancer cells and stromal cells according to breast cancer molecular
effects: on the origin of cancer cell energy metabolism and of yeast subtype. Breast Cancer Res 2013; 15: R78.
glucose repression. Biochim Biophys Acta 2011; 1807: 568–576. 45. Kim MJ, Kim DH, Jung WH, et al. Expression of metabolism-related
26. Swietach P, Wigfield S, Cobden P, et al. Tumor-associated proteins in triple-negative breast cancer. Int J Clin Exp Pathol 2014;
carbonic anhydrase 9 spatially coordinates intracellular pH in 7: 301–312.
three-dimensional multicellular growths. J Biol Chem 2008; 283: 46. Sotgia F, Whitaker-Menezes D, Martinez-Outschoorn UE, et al.
20473–20483. Mitochondrial metabolism in cancer metastasis: visualizing tumor
27. Casciari JJ, Sotirchos SV, Sutherland RM. Variations in tumor cell mitochondria and the ’reverse Warburg effect’ in positive lymph
cell growth rates and metabolism with oxygen concentration, glu- node tissue. Cell Cycle 2012; 11: 1445–1454.
cose concentration, and extracellular pH. J Cell Physiol 1992; 151: 47. Doyen J, Trastour C, Ettore F, et al. Expression of the
386–394. hypoxia-inducible monocarboxylate transporter MCT4 is increased
28. Schafer ZT, Grassian AR, Song L, et al. Antioxidant and oncogene in triple negative breast cancer and correlates independently with
rescue of metabolic defects caused by loss of matrix attachment. clinical outcome. Biochem Biophys Res Commun 2014; 451:
Nature 2009; 461: 109–113. 54–61.
29. Ullah MS, Davies AJ, Halestrap AP. The plasma membrane lac- 48. Witkiewicz AK, Whitaker-Menezes D, Dasgupta A, et al. Using the
tate transporter MCT4, but not MCT1, is up-regulated by hypoxia ’reverse Warburg effect’ to identify high-risk breast cancer patients:
through a HIF-1α-dependent mechanism. J Biol Chem 2006; 281: stromal MCT4 predicts poor clinical outcome in triple-negative
9030–9037. breast cancers. Cell Cycle 2012; 11: 1108–1117.

Copyright © 2015 Pathological Society of Great Britain and Ireland. J Pathol 2015; 237: 152–165
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com
MCT4 supports breast cancer cell survival 165

SUPPLEMENTARY MATERIAL ON THE INTERNET


The following supplementary material may be found in the online version of this article:
Supplementary materials and methods
Figure S1 (related to Figure 1). Panel of breast epithelial cell lines
Figure S2 (related to Figure 1). Validation of MCT4 as a differential target in breast cancer
Figure S3 (related to Figure 3). Metabolic characterization of MCT4-depleted breast cancer cells
Figure S4 (related to Figure 5). Analysis of Akt phosphorylation in breast epithelial cell lines
Figure S5 (related to supplementary methods). siRNA screen quality control and reproducibility
Table S1 (related to Figure 1). Results of siRNA screen targeting 231 metabolic genes
Table S2 (related to Figure 1). Validation of screen results
Table S3 (related to Figure 3). Metabolite uptake and secretion rates after MCT4 depletion

Copyright © 2015 Pathological Society of Great Britain and Ireland. J Pathol 2015; 237: 152–165
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com

You might also like