ArticlePDF Available

Epidermal growth factor-induced hepatocellular carcinoma: Gene expression profiles in precursor lesions, early stage and solitary tumours

Authors:

Abstract and Figures

Epidermal growth factor is an important mitogen for hepatocytes. Its overexpression promotes hepatocellular carcinogenesis. To identify the network of genes regulated through EGF, we investigated the liver transcriptome during various stages of hepatocarcinogenesis in EGF2B transgenic mice. Targeted overexpression of IgEGF induced distinct hepatocellular lesions and eventually solid tumours at the age of 6-8 months, as evidenced by histopathology. We used the murine MG U74Av2 oligonucleotide microarrays to identify transcript signatures in 12 tumours of small (n=5, pooled), medium (n=4) and large sizes (n=3), and compared the findings with three nontumorous transgenic livers and four control livers. Global gene expression analysis at successive stages of carcinogenesis revealed hallmarks linked to tumour size. A comparison of gene expression profiles of nontumorous transgenic liver versus control liver provided insight into the initial events predisposing liver cells to malignant transformation, and we found overexpression of c-fos, eps-15, TGIF, IGFBP1, Alcam, ets-2 and repression of Gas-1 as distinct events. Further, when gene expression profiles of small manifested tumours were compared with nontumorous transgenic liver, additional changes were obvious and included overexpression of junB, Id-1, minopontin, villin, claudin-7, RR M2, p34cdc2, cyclinD1 and cyclinB1 among others. These genes are therefore strongly associated with tumour formation. Our study provided new information on the tumour stage-dependent network of EGF-regulated genes, and we identified candidate genes linked to tumorigenes and progression of disease.
Content may be subject to copyright.
ORIGINAL PAPERS
Epidermal growth factor-induced hepatocellular carcinoma: gene
expression profiles in precursor lesions, early stage and solitary tumours
Ju
¨
rgen Borlak*
,1,4
, Tatiana Meier
1,4
, Roman Halter
1,4
, Reinhard Spanel
2
and Katharina Spanel-Borowski
3
1
Department of Pharmacology and Molecular Medicine, Fraunhofer Institute of Toxicology and Experimental Medicine, Nikolai-
Fuchsstr. 1, 30625 Hannover, Germany;
2
Institute of Pathology, Viersen, Germany;
3
Institute of Anatomy, University of Leipzig,
Germany
Epidermal growth factor is an important mitogen for
hepatocytes. Its overexpression promotes hepatocellular
carcinogenesis. To identify the network of genes regulated
through EGF, we investigated the liver transcriptome
during various stages of hepatocarcinogenesis in EGF2B
transgenic mice. Targeted overexpression of IgEGF
induced distinct hepatocellular lesions and eventually solid
tumours at the age of 6–8 months, as evidenced by
histopathology. We used the murine MG U74Av2
oligonucleotide microarrays to identify transcript signa-
tures in 12 tumours of small (n ¼ 5, pooled), medium
(n ¼ 4) and large sizes (n ¼ 3), and compared the findings
with three nontumorous transgenic livers and four control
livers. Global gene expression analysis at successive stages
of carcinogenesis revealed hallmarks linked to tumour size.
A comparison of gene expression profiles of nontumorous
transgenic liver versus control liver provided insight into
the initial events predisposing liver cells to malignant
transformation, and we found overexpression of c-fos, eps-
15, TGIF, IGFBP1, Alcam, ets-2 and repression of Gas-1
as distinct events. Further, when gene expression profiles of
small manifested tumours were compared with nontumor-
ous transgenic liver, additional changes were obvious and
included overexpression of junB, Id-1, minopontin, villin,
claudin-7, RR M2, p34cdc2, cyclinD1 and cyclinB1 among
others. These genes are therefore strongly associated with
tumour formation. Our study provided new information on
the tumour stage-dependent network of EGF-regulated
genes, and we identified candidate genes linked to tumori-
genes and progression of disease.
Oncogene (2005) 24, 1809–1819. doi:10.1038/sj.onc.1208196
Published online 24 January 2005
Keywords: HCC; EGF; transgenic mice; tumour stages;
gene expression profiling
Introduction
It is estimated that about 350 000 new cases of
hepatocellular carcinoma (HCC) arise per year (Schafer
and Sorrell, 1999). The most prominent risk factors are
chronic hepatitis B and C virus infection, Aflatoxin B1
exposure and alcohol-related cirrhosis. As of today, the
precise molecular mechanism in the onset and progres-
sion of disease remains uncertain.
Overexpression of liver mitogens may be an important
mechanism of disease. EGF and TGFa are potent
mitogens for hepatocytes (Wang et al., 1999; Rescan
et al., 2001). Signalling of these mitogens is through
binding to members of the EGF-receptor family. Their
expression is unregulated in HCC and this supports
autocrine growth stimulation of hepatoma cells (Yama-
guchi et al., 1995; Chung et al., 2000). EGF also plays an
important role in hepatocyte morphology (Rescan et al.,
2001). Overexpression of EGF might be an important
step towards development of liver cancer and is
suspected to play a particular role in spontaneous liver
tumour development (Ostrowski et al., 2000). There is
suspicion that EGF plays a role in Helicobacter
hepaticus-induced chronic hepatitis with progression to
hepatocellular cancer (Ramljak et al., 1998). Previous
investigations demonstrated targeted overexpression of
a secretable form of EGF (IgEGF) to result in multiple
highly malignant HCCs, with 100% fatalities around
7–8 months after birth (To
¨
njes et al., 1995). This
transgenic mouse line therefore mimics effectively the
consequence of altered EGF signalling via the EGF
receptor. We used this mouse model to identify the
network of EGF-regulated genes at various stages of
tumour development and in solid tumours. Global gene
expression analysis at successive stages of carcinogenesis
holds promise for an identification of master genes for
the onset and progression of disease. Thus, we aimed to
identify candidate genes associated with early stages of
tumorigenesis and with developed HCCs. Overall, this
study aimed for a better understanding of the network
of EGF-regulated genes in liver carcinogenesis.
Results
Histopathology of liver tumours
Macroscopically, all EGF-overexpressing animals devel-
oped tumours at the age of 6–9 months. A total of six
Received 20 February 2004; revised 28 July 2004; accepted 9 August
2004; published online 24 January 2005
*Correspondence: J Borlak; E-mail: Borlak@item.fraunhofer.de
4
These authors contributed equally to this work
Oncogene (2005) 24, 1809–1819
&
2005 Nature Publishing Group
All rights reserved 0950-9232/05 $30.00
www.nature.com/onc
tumour-bearing EGF2B mice as well as tumour-free
parenchyma and liver from wild-type animals were
investigated. The HCCs ranged from less to highly
differentiated. As a rule, small HCCs were adenoma-like
tumours with well-developed hepatocytes, whereas
larger HCC showed a lot of cellular dedifferentiation
and enhanced nuclear atypia (Figure 1).
There were precursor lesions in the tumour-free liver
tissue as well. Basically all of the trabecular parenchyma
showed nuclear atypia with enlarged nuclei, mainly
increased numbers of large polyploid nuclei, anisocar-
yosis and some polymorphism, defined as large-cell
dysplasia (LCD). This LCD seemed to merge into multi-
centric nodule formation, usually with small cell changes
and various degrees of atypia. According to the actually
proposed terminology of nodular hepatocellular lesions
in human pathology (Hepatology 1995, International
Working Party), these nodules can be defined as
dysplastic foci (DF) and dysplastic nodules (DN).
Transcript profiling
Abundant expression of transgenic EGF in liver and
tumours of transgenic mice was confirmed by RT–PCR,
as depicted in Figure 2. As transcript profiling of
juvenile tumours is of considerable value for an
identification of priming factors in tumour development,
we divided tumours into three groups according to their
size. We investigated gene expression profiles in devel-
oping liver carcinomas and compared global expression
profiles of wild-type animals with liver tumours from
small to large size, as well as macroscopically non-
tumorous livers of tumour-bearing animals. In all, 4175
genes and ESTs were commonly expressed in all eight
tumour samples and 4149 genes were expressed in n ¼ 4
normal liver samples, the expression of many genes
being increased or decreased during the process of liver
carcinogenesis (Table 1). To determine significant gene
expression changes, we performed T-test analysis
between control liver and sets of tumours, taking only
changes into account, which were detected in all tumour
samples or in all control livers for identification of
overexpressed or repressed genes accordingly. Further,
we performed ranking analyses (see Materials and
methods) that enabled a more stringent comparison
according to the consistency of gene expression changes
(Table 1). Notably, gene expression profiles were
significantly changed in transgenic nontumorous liver,
but the number of deregulated genes was further
increased in tumours. In all, 109 of the upregulated
and 55 of the repressed genes show a signal intensity
X70, a fold change X3orp3, a P-value p0.05 and
100% concordance of expression changes in pairwise
comparative analyses in at least one group of the
tumours, when compared with normal liver. A total of
59 of the overexpressed and five of the repressed genes
selected by their importance and their possible functions
are listed in Table 2a and b.
Some genes known to be important for cancer biology
(e.g. cyclinD1 and PDGFa) with a lower FC are also
included. For the pool of small tumours, T-test analysis
could not be carried out. Instead, the inclusion criterion
was 100% concordance of expression changes, when
compared to n ¼ 4 normal liver samples. Table 2
allows a direct comparison of gene expression profiles
of different tumour sizes, from EGF-overexpressing
Figure 1 Histology of tumours and precursor lesions in EGF2B transgenic mice. Normal liver tissue of nontransgenic controls with
normal polyploidal variation of hepatocytic nuclear size (central vein: cv) (a). LCD in tumour-free parenchyma of EGF2B mice
merging with initial perivenous DF (arrows) and DN (arrowheads): Small-cell nodular proliferations (b). Well-differentiated,
adenoma-like small HCC: irregular, mainly bilayered trabeculae, slight polymorphism and lipid vacuolization (c). Multilayered
trabeculae of larger HCC (d). Same magnifications. Staining: H (a, b); PAS (c); H and E (d)
Early tumour stage in EGF-induced HCC
J Borlak et al
1810
Oncogene
tumour-free transgenic liver with LCD to large mani-
fested tumours. We found certain genes to be equally
deregulated in displastic transgenic liver as in tumours
(e.g. c-fos, eps-15, EGR-1, TGIF, IGFBP1, Alcam),
while others were dramatically deregulated at the onset
of small HCC (e.g. p34cdc2, Id1, junB, minopontin,
claudin7). Besides, we identified 23 genes that were
uniquely expressed in all tumours, but not in controls
Table 1 Number of genes and ESTs expressed in EGF-induced liver tumours and normal liver
Genes or ESTs EGF-transgenic liver
(three samples)
All tumours
(eight samples)
Normal liver
(four samples)
Detected in all samples within a group 4494 4175 4149
Not detected in all samples within a group 6569 5768 6655
Upregulated in liver tumours according to T-test (P-valuep0.05); FCX2 140 265
Upregulated in liver tumours according to comparison ranking (100%
concordance in comparative analyses); FCX2
67 89
Downregulated in liver tumours according to T-test (P-valuep0,05); FCp2 52 130
Downregulated in liver tumours according to comparison ranking (100%
concordance in comparative analyses); FCp2
23 59
Uniquely expressed either in all tumours or normal liver, P-value in T-test
p0.05, 100% concordance in comparative analyses; FCX2/p2
23 1
Figure 2 (a) Structure of the transgene. albP, murine albumin promoter; Ig-S, Ig-signal sequence; I, Intron sequence; EGF, synthetic
EGF; SVA, SV40 poly-A signal. (b) PCR analysis of Ig/EGF from tail biopsies to identify transgenic mice. Lane 1: nontransgenic mice;
lanes 2–5: transgenic mice; M: molecular weight standard; lane 6: amplified fragment of the transgene was digested with EcoRI to
obtain fragments of 210 and 107 bp. (c) RT–PCR analysis of the housekeeping gene (b-actin) and the transgene. Lanes 1–3:
nontransgenic liver, lanes 4–6: macroscopic nontumorous liver of transgenic mice, lane 7: pool of small tumours; lanes 8–10: tumours
of medium size; lanes 11 and 12: tumours of large size; lane 13: negative control (water). (d) RT–PCR of selected genes (lanes represent
samples as described in (c))
Early tumour stage in EGF-induced HCC
J Borlak et al
1811
Oncogene
Table 2 Gene expression signatures in EGF-induced mouse liver tumours: (a) upregulated genes; (b) downregulated genes
Gene ACC Gene description Transgenic nontumorous
liver
Tumours
Small size (o2 mm)
(Pool)
Medium size (5 mm) Large size (10 mm)
FC P-value
(%)
FC %
a
FC P-value
(%)
FC P-value
(%)
(a) Upregulated in EGF-induced liver tumours
b
Upregulated in transgenic liver and tumours
Unknown function
Trefoil factor 3 D38410 Mucin-associated polypeptide 105.8 0.009 164.3 100 117.8 0.156 (75) 50.1 0.372 (66)
H19mRNA X58196 Post-transcriptional regulator,
overexpressed in HCC
24.84 0.317 164.8 100 136.1 0.046 40.55 0.269
CD63 antigen D16432 Tetraspanin, late endosome
marker
11.33 0.009 34.98 100 27.65 0.018 50.35 0.005
G7e U69488 Viral envelope-like protein 7.65 0.052 (83) 34.39 100 19.91 0.005 23.96 0.007
Fibrinogen-like protein 2 M16238 Unknown 6.7 0.081 5.01 100 6.52 0 8.26 0.006
Growth promotion
PDGFa M29464 Platelet-derived growth factor a,
induced in HCC
2.05 0.011 (41) 2.77 100 2.51 0.022 (87) 2.44 0 (83)
TGFa M92420 Transforming growth factor a,
induced in HCC
1.69 0.027 (41) 1.93 100 2.07 0.026 (68) 3.62 0.036
eps-15 L21768 Substrate of the EGF-receptor,
transforming capacity
2.82 0.02 2.89 100 2.48 0.005 (50) 3.44 0.012
FBJ osteosarcoma
oncogene (c-fos)
V00727 Oncogene, signal transduction 2.82 0 4.09 100 2.9 0.048 5.64 0.18
Insulin-like growth
factor-binding protein I
(IGFBP1)
X81579 Regulation of cell growth 18.8 0.186 (75) 21.18 100 20.83 0.032 30.48 0.005
TGIF X89749 Inhibitor of TGFb signalling 2.79 0.107 2.56 100 2.6 0.011 3.62 0.067
Transcription factor
ets-2 J04103 Transcription factor, induced in
HCC
3.94 0.102 (91) 3.44 100 3.61 0.002 4.25 0
Angiogenesis
Early growth response
(EGR-1)
M28845 Zinc-finger encoding gene,
regulates tumour angiogenesis
10.9 0.103 (91) 9.06 100 6.65 0.007 (93) 8.26 0.021
Adhesion
Alcam L25274 Adhesion molecule, involved in
tumour development
3.84 0.006 5.18 100 3.89 0.032 4.69 0.008
Defense response
TIS21 M64292 Negative control of cell growth 6.46 0.111 (66) 5.64 100 5.18 0.093 (25) 5.92 0.002 (83)
B-cell translocation gene 3
(ANA)
D83745 Antiproliferative protein 7.17 0.093 8.48 100 10.15 0.001 22.42 0.074
BLNK AF068182 Central linker protein in B-cell
activation
2.25 0 3.65 100 3.48 0.036 4.54 0.011
Metabolism
Stearoyl-CoA desaturase 2 M26270 Fatty acid biosynthesis;
upregulated in HCC
12.13 0.062 33.4 100 34.91 0.062 28.62 0.01
Lipoprotein lipase M63335 Fatty acid degradation 27.02 0.033 22.25 100 20.03 0.019 23.78 0.05
Early tumour stage in EGF-induced HCC
J Borlak et al
1812
Oncogene
Table 2 (Continued )
Gene ACC Gene description Transgenic nontumorous
liver
Tumours
Small size (o2 mm)
(Pool)
Medium size (5 mm) Large size (10 mm)
FC P-value
(%)
FC %
a
FC P-value
(%)
FC P-value
(%)
Retinol-binding protein I X60367 Vitamin A, E binding 7.87 0.005 13.29 100 10.03 0.001 9.99 0.009
Extracellular matrix
Nidogen1 L17324 Basement membrane compo-
nent, cell–matrix interaction, cell
adhesion
1.95 0.002 (100) 3.05 100 2.26 0 (100) 3.66 0.018 (100)
Miscellaneous
Phospholipid scramblase 1
(TRA1)
D78354 Potential role in growth factor
signalling pathways; associated
with leukomogenesis
3.61 0.083 4.87 100 4.86 0.003 5.65 0.016
Overexpressed at the onset of small tumours
Cell cycle promotion
CyclinB1 X64713 Cell cycle control Absent 2.85 100 1.92 0.007 (68) 3.07 0.036
CyclinD1 AI849928 Cell cycle control 1.19 (100) 2.79 100 2.88 0.019 (87) 2.68 0.119 (83)
Cell division cycle control
protein 2a (p34 cdc2)
M38724 Mitosis-specific phosphory-
lation of cytoskeletal protein
Absent 5.74 100 4.82 0.026 (0) 6.77 0.05 (66)
Ki67 X82786 Tumour cell proliferation
marker
1.28 0.004 (66) 3.88 100 2.64 0.008 3.55 0.074 (83)
Transcription factors
Inhibitor of DNA binding
1 (Id-1)
M31885 Helix–loop–helix protein;
inactivates p16/pRB pathway in
prostate cancer
Absent 4.41 100 1.57 0.020 (12) 2.98 0.013 (66)
junB U20735 Transcription factor,
proto-oncogene
Absent 3.38 100 1.97 0.232 (0) 2.77 0.091 (58)
Ets transcription factor
(ELF3)
AF016294 Member of ETS family,
involved in cancer
1.84 0.202 (16) 3.19 100 2.88 0.011 (25) 2.22 0.047 (25)
Signalling
Cell adhesion kinase L57509 CAK receptor kinase Absent 7.31 100 5.03 0.006 (25) 9.2 0.013
Cytoskeleton
villin M98454 Matrix protein in microtubuli;
upregulated HCC
Absent 8.75 100 1.45 0.166 (25) 1.87 0.178 (66)
Angiogenesis
Calpactin I (annexin II) M14044 Calcium-dependent
phospholipid binding;
upregulated in proliferating
hepatocytes; role in angiogenesis
1.92 0.019 (75) 7.86 100 7.6 0.018 19.17 0.07
Lymphotoxin b U16985 TNF family cytokine, lymph
node development; may initiate
angiogenesis
Absent 23.58 100 5.73 0.01 (75) 13.19 0.276
Adhesion
claudin-7 AF087825 Tight junction adhesion
protein, activates processing of
pro-matrix metalloproteinase-2
Absent 4.69 100 Absent Absent
Cell death
Cell death factor CIDE-A AF041376 DNA fragmentation, activation
of apoptosis
Absent 5.92 100 3.94 0.068 (50) 4.48 0.225 (33)
Early tumour stage in EGF-induced HCC
J Borlak et al
1813
Oncogene
Table 2 (Continued )
Gene ACC Gene description Transgenic nontumorous
liver
Tumours
Small size (o2 mm)
(Pool)
Medium size (5 mm) Large size (10 mm)
FC P-value
(%)
FC %
a
FC P-value
(%)
FC P-value
(%)
Invasion, metastasis
Minopontin (osteopontin) X13986 Secreted adhesive glycoprotein,
upregulated in HCC
1.26 0.422 (0) 18.22 100 2.01 0.593 (25) 3.39 0.067
Ribonucleotide reductase
M2 subunit (RR M2)
M14223 DNA synthesis and repair, role
in cancer/metastases
1.35 0.046 (0) 4.49 100 2.8 0.024 (75) 2.06 0.001 (8)
Lymphocyte antigen 6
complex (Ly6d)
X63782 GPI-anchored protein,
correlates with malignancy
of mouse tumours
Absent 10.67 100 17.29 0.041 47.42 0.177
Proteolysis, peptidolysis
CarboxypeptidaseE X61232 Abundantly expressed in
hepatoma and HCC
0.6 0.253 (58) 3.71 100 5.02 0.195 52.71 0.162
Metabolism
Nonallelic mRNA for
pancreatic a-amylase
isoenzyme (pCEPa12)
X02578 Overexpressed in lung cancers 1.67 0.150 (60) 63.51 100 47.27 0.287 (75) 85.87 0.14
Transport
Rab3D M89777 Small GTPase, regulatory role
in vesicular transport
Absent 14.73 100 8.16 0.002 13.87 0.056
Solute carrier family 7 AJ012754 Cationic amino-acid
transporter
Absent 6.05 100 7.03 0.001 5.13 0.001 (66)
Immune response
Toll-like receptor 6
(TLR6)
AB02088 Activation of Nf-kB and c-Jun
N-terminal kinase (JNK),
immune response
2.22 0.002 (58) 7.19 100 3.75 0.013 (81) 3.67 0.015
Miscellaneous
Endothelial monocyte
activated polypeptideI
(EMAP)
U41341 Tumour-derived cytokine 3.45 0.048 (33) 28.63 100 17.84 0.04 59.61 0.082
Other genes upregulated in tumours
Signalling
Ect2 oncogene L11316 Signal transduction,
Rho-specific exchange factor
2.19 0.088 (58) 3.96 100 2.83 0.01 (93) 3.55 0.088 (83)
A6 related protein Y17808 Mouse homolog of human
protein kinase
2.4 0.144 (50) 6.42 100 6.98 0.013 14 0.013
Transcription factor
LRG-21 U19118 Transcription factor 3.1 0.179 (33) 4.65 100 4.05 0.119 5.49 0.015
Angiogenesis, invasion, metastasis
RhoC X80638 Small GTPase, involved in
angiogenesis and metastasis
2.17 0.026 (66) 4.11 100 3.11 0.01 7.45 0
Plasminogen activator
inhibitor-1 (PAI-1)
M33960 Serine protease inhibitor
involved in angiogenesis,
invasion, metastasis
4.43 0.187 (66) 1.21 0 3.35 0.273 (25) 12 0.01
Serpinb6 U25844 Serine protease inhibitor 3,
regulation of tumour
progression, inflammation
and cell death
1.29 0.077 (41) 1.81 100 2.34 0.082 (81) 8.06 0.007
Early tumour stage in EGF-induced HCC
J Borlak et al
1814
Oncogene
Table 2 (Continued )
Gene ACC Gene description Transgenic nontumorous
liver
Tumours
Small size (o2 mm)
(Pool)
Medium size (5 mm) Large size (10 mm)
FC P-value
(%)
FC %
a
FC P-value
(%)
FC P-value
(%)
Extracellular matrix
Procollagen type IV a 1 M15832 Basement membrane-associated
protein, cell adhesion, upregu-
lated in human HCC
1.66 0.083 (58) 3.27 100 3.94 0.006 (100) 3.81 0.204 (100)
Cytoskeleton
Cytokeratin endoA X15662 Intermediate filament protein,
specific to carcinoma
2.01 0.203 (66) 3.14 100 2.4 0 (100) 4.48 0.081 (100)
Cytoceratin endoB M22832 Intermediate filament protein,
specific to carcinoma
2.39 0.098 (83) 3.83 100 2.73 0 (100) 4.74 0.041 (100)
Miscellaneous
AE binding protein AF053943 Wound healing 3.94 0.169 (33) 8.43 100 7.81 0 16.5 0.032
Reduced expression 3
(REX-3)
AF05134 Downregulated by retinoic acid-
induced growth inhibition in
murine teratocarcinoma cells
1.46 0.031 9.03 100 4.45 0 14.42 0.035
Carbonic anhydrase 2 M25944 Role in facilitating transport of
CO
2
, regulated by micro-
environmental hypoxia
2.22 0.013 (41) 5.26 100 8.0 0.122 (81) 22.79 0.226 (66)
MT-ACT48 AJ238894 Mitochondrial long-chain
acyl-CoA thioesterase
(eps-15 partner)
1.36 0.134 (16) 2.05 75 1.66 0.003 (62) 3.4 0.001
ww domain-binding 5 U92454 Binding domain for proline-rich
sequences of various structural,
regulatory and
signalling proteins
1.49 0.034 4.5 100 3.68 0 (57) 5.18 0.036
rbm3 AB016424 Cold shock protein 2.24 0.009 (75) 5.36 100 3.47 0.029 (93) 7.36 0.109
(b) Downregulated in EGF-induced liver tumours
c
Cyt. P450 retinoic acid Y12657 RA oxidation 3.3 0.003 (91) 5.43 100 4.23 0.001 5.05 0.004
Growth arrest-specific 1
(GAS-1)
X65128 Cell proliferation inhibitor 2.42 0.005 (75) 3.98 100 4.81 0.002 6.69 0.002
Tyrosinase-related protein-2
(tyrp-2)
X63349 Melanine biosynthesis 2.77 0.059 (58) 13.27 75 11.3 0.002 (93) 24.85 0.018
Cytochrome P450 2f2 M77497 Naphthalene detoxification 1.36 0.285 (8) 10.51 100 7.24 0.006 12.79 0.006
SULT-X1 AF02604 Sulphotransferase-related
protein
1.7 0.089 (50) 16.27 100 8.67 0.01 16.27 0.009
ACC: accession number; FC: fold change; P-value: P-value in T-test; %: Concordance (%) of change calls in the pairwise comparisons (each tumour compared to each normal liver sample) by
which genes were up- or downregulated. If not indicated, concordance ¼ 100%.
a
Instead of P-value for the pool of small tumours, the concordance (%) is shown, because for a pool no T-test
analyses could be carried out.
b
The genes had in at least one group of the tumours an FCX3, a P-value in T-test o0.05, a signal intensity >70 and were upregulated in 100% of pairwise analyses,
compared with normal liver of control mice. Grey-coloured genes were absent in the control tissues. Grey-coloured rows indicate genes expressed in all tumours and not in controls.
c
The genes had
in at least one group of tumours an FC o3, a P-value in T-test o0.05, were downregulated in pairwise analyses when compared to normal liver of control mice and a signal intensity >70 in
control liver. Grey-coloured rows are genes, which were expressed in all normal tissues but not in tumours.
Early tumour stage in EGF-induced HCC
J Borlak et al
1815
Oncogene
(Table 1), and these included Ets transcription factor,
calpactin, stearoyl-CoA desaturase 2 and Rab3D among
others (Table 2). In contrast, Gas1 and tyrosinase-
related protein-2 were detected in all normal liver
samples, but not in tumours (Tables 1, 2).
Notably, among genes which were upregulated in
tumours (Table 2a), we found several genes with
an inferred role in tumour development such as
growth factors TGFa, some components of EGF/
TGFa-mediated signalling pathway (c-fos, eps-15, MT-
Act48), EGR-1, cell division cycle control protein 2a,
proto-oncogene junB, IGFBP1, cyclinB1, cyclinD1 and
Id-1, the latter two interfering with the Rb pathway.
Additionally, genes that are associated in human cancers
with metastasis and angiogenesis like rhoC, PAI-1 and
calpactin 1 (annexin II) were upregulated. Further, we
observed overexpression of some genes coding for
components of the cytoskeletal network (villin, cytoker-
atin endoA and B) and basal membrane (collagen IVa1,
nidogen), and proteins that are involved in survival
factor signalling pathway like IGFBP1 and TLR6. The
most highly overexpressed genes in all tumours included
H19 mRNA, trefoil factor 3, stearoyl-CoA desaturase 2,
lipoprotein lipase, nonallelic mRNA for pancreatic
a-amylase isoenzyme, CD63 antigen, G7e their
contribution in carcinogenesis may be of great impor-
tance, but requires further study. The expression of
certain genes was obviously enhanced as part of a
defence response to the free development, for example,
cell death factor CIDE-A, TIS21 and ANA (BTG3). In
strong contrast, most of the repressed genes code for
proteins important for liver function, including drug
binding, detoxification and metabolism (see Table 2b).
We thus provide evidence for loss of metabolic
competence in liver tumour tissue.
The expression of a selected number of genes was
additionally analysed by semiquantitative RT–PCR
(Figure 2d). There was good agreement between
microarrays and RT–PCR experiments, for example,
no expression of G7e and Trefoil factor 3 in controls
but strong expression of TGIF and LRG-21 in large
tumours and strong expression of Trefoil factor 3 and
G7e in small tumours. In the case of calpactin, the level
of induction determined by RT–PCR and the micro-
array differed, though an identical trend was observed,
for example, induction. Presumably, the different
methods of gene expression analysis (see Materials and
methods) and the algorithm applied for data analysis of
microarray experiments produced different estimate of
induction levels.
Discussion
Liver pathology
Morphological phenotyping enabled new insights into
the process of hepatocellular carcinogenesis. Particu-
larly, the onset and development of malignant tumours
were followed. Notably, no benign tumours were found.
We assume LCD to be the precursor lesion being at the
edge of malignant change. No inflammatory process
was observed and no cirrhotic conversion of liver tissue
was noted. This contrasts with microscopic lesions
frequently seen in humans, where cirrhotically altered
organs are the main precursor of carcinogenesis
(Kubicka et al., 2000). An important finding of our
study was the 100% incidence of malignant tumour
within 6–8 months after birth, and we suggest the
following sequence of events: diffuse LCD merges into
multiple DF and DN, with local growth towards HCC.
Transcript profiling
The transgenic mouse line EGF2B develops HCCs
(To
¨
njes et al., 1995) as a consequence of overexpression
of a secretable form of EGF, which is known to be a
strong mitogen for liver cells. There is cumulative
evidence for EGF or TGFa overexpression to be
necessary, but not sufficient in inducing carcinogenesis
in mice (Sandgren et al., 1993; Wu et al., 1994; To
¨
njes
et al., 1995). As shown in our study and by others,
undue exposure to EGF predisposes liver tissue to
cancer. Large-scale expression analysis enabled initial
changes to be studied. We compared the expression
profiles of tumour-free liver of transgenic with normal
liver of control mice. In transgenic displastic liver,
certain genes were upregulated at the same level as in
tumours. These genes include eps-15, a substrate of
EGF-R with transforming capacity (Alvarez et al.,
1995), and c-fos, a transcription factor of the EGF
signalling pathway known to be induced by EGF and
other growth factors (Dey et al., 1991). Therefore,
transcriptional activation of the EGF/TGFa signalling
pathway in the liver and tumours of EGF2B transgenic
mice is a prominent feature. Similarly, the transcrip-
tion factor EGR-1, which is co-regulated with c-fos
(Chavrier et al., 1989), was also overexpressed. Recently,
EGR-1 was shown to play an important role in tumour
angiogenesis and growth (Fahmy et al., 2003). Remark-
ably, we observed a strong upregulation of insulin-like
growth factor-binding protein 1 (IGFBP1) (Table 2a),
a hepatocyte-derived and secreted protein, which is
required for liver regeneration. The recent report of Leu
et al. (2003) provided evidence for IGFBP1 to function
as a critical hepatic survival factor in the liver by
reducing the level of proapoptotic signals. Therefore,
overexpression of IGFBP1 may lead to enhanced
survival of tumorous liver cells.
Further, the role of BLNK (Table 2a), which is
involved in activation of nuclear factor NF-kB (Tan
et al., 2001) and Toll-like receptor, which activates both
NF-kB and c-jun N-terminal kinase (JNK) (Takeuchi
et al., 1999), is difficult to comprehend, because these
genes are mainly induced in immune-competent cells
after stimulation. On the other hand, their expression
in the liver would indicate an imbalance of the IGF/
IGFBP system of EGF2B transgenic mice and support
the survival of tumorous liver cells. As no infiltration of
the tumour tissue by immune cells was observed,
enhanced expression of these genes may support
a function beyond immune-competent cells. There is
Early tumour stage in EGF-induced HCC
J Borlak et al
1816
Oncogene
a clear need to clarify the role of these genes in HCC
formation.
Further, elevated expression level of TGIF, an
inhibitor of antigrowth factor TGFb-responsive trans-
cription (Melhuish et al., 2001) and transcription factor
ets-2 among others (Table 2a), can also contribute to
the initial transformation of hepatocytes in EGF-
overexpressing mice.
Overexpression of an adhesion molecule Alcam, a
member of the immunoglobulin superfamily, may
contribute to the invasive capabilities of transformed
liver cells (Choi et al., 2000). Remarkably, two genes
involved in lipid metabolism, for example, lipoprotein
lipase and stearoyl-CoA desaturase-2, were uniquely
expressed or strongly induced in liver and tumours of
transgenic mice. This suggests that important functions
of the coded genes in lipid and fatty acid metabolism of
tumour cells most probably contribute to cell membrane
synthesis.
A comparison of the expression profiles of EGF-
overexpressing tumour-free transgenic liver with those
observed in manifested small liver tumours allows an
identification of candidate genes additionally required
for malignant transformation. We observed proto-
oncogene junB, cell division cycle control protein
p34cdc2, cyclinD1, cyclinB1, Id-1, minopontin, villin,
claudin-7, ribonucleotide reductase (RR M2), Ly6d and
cell adhesion kinase, which were not changed in
transgenic liver (mostly not detected both in control
and transgenic liver), but were dramatically induced in
the pool of small tumours (Table 2a). CyclinD1 forms a
complex with cdk4 to inactivate Rb by phosphorylation
and Id-1 was shown to inactivate the p16/pRB pathway
by preventing the deactivation of cyclin/cdk complexes
in human prostate cancer (Ouyang et al., 2002). Thus,
the overexpression of cyclinD1 and Id1 in liver cells
can interfere with pRb pathway, leading to exaggerated
cell division signalling. We observed overexpression of
minopontin (osteopontin), RRM2 and Ly6d in small
tumours of EGF2B mice. Overexpression of the latter
genes was observed in human cancers and correlated
with invasiveness and metastatic potential (Chen et al.,
2000; Witz, 2000; Gotoh et al., 2002).
Many regulated genes showed permanent increase
or decrease in their expression level during carcino-
genesis. Importantly, we failed to detect Gas-1 in all
tumour samples, while its expression was evident in all
controls and a low expression in two of three
nontumorous transgenic livers (see Table 2a). This
protein is of importance in growth suppression and
it was suggested that pRb and/or p53 play an active
role in mediating the growth-suppressor effect of Gas-1
(Del Sal et al., 1994, 1995; Evdokiou and Cowled, 1998).
It appears that loss of growth control through GAS-1
may be a necessary event in the multi-step neoplastic
transformation.
Additionally, we observed a significant increase in the
transcript level of the small GTPase rhoC (Table 2a),
and of Ect2 (Table 2a), the guanine nucleotide exchange
factor for Rho GTPases (Tatsumoto et al., 1999), which
plays a critical role in Rho activation (Kimura et al.,
2000). RhoC is involved in controlling cell motility and
focal adhesion, and was recently demonstrated to be
associated with vascular invasion in human HCC
(Okabe et al., 2001) and may play a role in metastasis
of human melanoma (Clark et al., 2000).
Transcript profiling of large-size tumours evidenced
induction of PAI-1, serpin b6, calpactin (annexin II),
carboxypeptidase E and EMAP. Their specific role in
tumour progression still needs to be delineated.
It is apparent that the EGF-transgenic mouse model
is valuable for the study of HCC. Indeed, the tumours in
these animals share known features with those pre-
viously observed in humans following EGF induction or
by malignant transformation, and our analyses revealed
novel candidate genes associated with tumorigenesis.
This included Rab 3D, cell adhesion kinase, trefoil
factor 3, A6-related protein, LRG-21, cold shock
protein cbm-3, AE-binding protein, ww domain-binding
protein, Tra1, fibrinogen-like protein and tyrosinase-
related protein-2, but their specific role in liver
carcinogenesis needs to be elucidated.
In conclusion, we report tumour size-dependent gene
expression in EGF-induced HCCs. We observed en-
hanced expression of villin, cell death factor CIDE-A,
claudin 7 and junB in specifically small tumours. We
further observed induction of autocrine growth with
increased expression of TGFa, PDGFa and eps-15,
the latter being a substrate for the EGF receptor with
transforming capacity. In all tumours, induction of c-fos
and egr-1 was significant, as was the induction of the
survival factor IGFBP1, which provided tumours with
an important advantage. In large tumours, RhoC
activation was linked to vascular invasion. Finally, loss
of sensitivity to antigrowth signals could be traced back
to induction of TGIF, Id-1 and cyclinD1 to interfere
with the pRb control of cell division, whereas the
repression of the tumour suppressor gene Gas-1 allowed
for proliferation invasion and metastatic growth. In
future, promoter analyses of deregulated genes is needed
to identify the molecular rules of promoter activation
and the transcription factors acting in concert in
malignancies of the liver. Most certainly, the EGF
transgenic mouse model contributes towards a mole-
cular understanding of liver carcinogenesis
Materials and methods
Maintenance of the transgenic mouse line
The EGF2B transgenic line was described earlier by To
¨
njes
et al. (1995). Transgenic mice were maintained as hemizygotes
in the CD2F1-(DBA/2 Balb/c) background. PCR was
carried out with Platinum PCRSuperMix (InVitrogen). An-
nealing temperature and the number of cycles are indicated in
brackets after each primer pair. The transgene was verified by
PCR of DNA extracted from tail biopsies (Hogan et al., 1994)
and the following forward primer (fp) and reverse primer
(rp) pair was used for a transgene-specific amplification:
forward primer: 5
0
-CTAGGCCAAGGGCCTTGGGGGCTC
TTGCAG-3
0
; reverse primer: 5
0
-CATGCGTATTTGTCCAG
AGCTTCGATGTA-3
0
(611C, 32 cycles).
Early tumour stage in EGF-induced HCC
J Borlak et al
1817
Oncogene
Gene expression studies by RT–PCR
RT–PCR was employed to confirm expression of the transgene
and some selected genes in controls, tumours and nontumor-
ous liver of transgenic mice. The primer design was carried out
with the MacVectort 6.5.3 software and cross-reaction of
primers with other genes was excluded by comparison of the
sequence of interest with a data bank (Blast 2.0 US National
Centre for Biotechnology information). We also used UCSC
genome bioinformatics to design intron-spanning primers.
Total RNA was isolated with the Qiagen RNA purification kit
according to the manufacturer’s instructions. Reverse tran-
scription was carried out using Omniscript (Qiagen), Oligo-dT
primers (InVitrogen) and RNasin (Promega), followed by
PCR amplification (see above) with the following primer pairs:
EGF, fp: GCTGTGACGGTCCTTACAATG; rp: CAGTTC
CCACCACTTCAGGTC (611, 29 cycles). Calpactin, fp: GAG
CATCAAGAAAGAGGTCAAAGG; rp: TTCAGTCATCC
CCACCACACAG (651C, 28 cycles). LRG21, fp: AGATGAG
AGGAAAAGGAGGCGG; rp: GGGTGGAAAAGGAGGA
TTCAGTAAG (651C, 28 cycles). G7e, fp: GGTCTTTCACA
AGCAGTGCCTG; rp: AAACCAAGTTCCAATGGGGG
(571C, 32 cycles). TFF3, fp: GCAAATGTCAGAGTGGACT
GTGG; rp: GGCTGTGAGGTCTTTATTCTTCAGG (621C,
28 cycles). TGIF, fp: AACGCCTATCCCTCAGAGCAAG;
rp: GTCCAACTACGCAGGAATGAAATG (651C, 30 cycles).
b-Actin was used as a housekeeping gene, because its
expression was found to be unchanged in controls (nontrans-
genic), transgenic (nontumorous) and tumours of EGF2B mice
(microarray analyses) The following primer pair was used: fp:
GGCATTGTTACCAACTGGGACG; rp: CTCTTTGATGT
CACGCACGATTTC (651C, 25 cycles). PCR reaction pro-
ducts were separated on 1% agarose gels stained with ethidium
bromide and photographed on a transilluminator (Kodak
1544 CF). A semiquantitative measurement was carried out
using Kodak 1D software (v.3.5.3)
Histology
Tumour tissue and tumour-free tissue of transgenic animals, as
well as liver from control animals, were fixed in 4%
formaldehyde in PBS and embedded in paraffin by standard
operating procedures. Paraffin blocks were sectioned into 3–
5 mm thick slices and stained with haematoxylin and eosin (H
and E), haematoxylin only (H) and PAS for light microscopic
evaluation.
Sample collection and preparation
Mice were anaesthesized by an overdose of CO
2
at the age of
6.5–9 months. The thorax was opened by standard surgical
procedures and the liver was explanted and rinsed with PBS.
The tumours were inspected macroscopically and separated
from the liver. One group consisted of four analysed tumours
of about 5 mm in size (medium size) derived from n ¼ 4
animals, and a second group of three tumours of about 10–
15 mm (large size) were derived from n ¼ 3 animals. Five
tumours from n ¼ 3 animals were about 1 mm (small size) and
were pooled to improve the yield in RNA. Tumour-free tissue
was taken from the liver of (n ¼ 3) tumour-bearing transgenic
mice. Healthy liver from n ¼ 4 non transgenic CD2F1 mice of
about the same age were used as controls. Upon anatomical
preparation tissue was frozen immediately in liquid nitrogen.
RNA isolation and production of copy RNA
The cRNA samples were prepared following the Affymetrix
Gene Chip
s
Expression Analysis Technical Manual (Santa
Clara, CA, USA). Briefly, total RNA was isolated from frozen
tissues using QIAGEN’s RNeasy total RNA isolation
procedure. A second cleanup of isolated RNA was performed
using the same RNA isolation kit. In all, 10 mg of total
RNA was used for the synthesis of double-stranded cDNA
with Superscript II RT and other reagents from Invitrogen
Life Technologies. HPLC-purified T7-(dT)
24
(GenSet SA)
was used as a primer. After cleanup, double-stranded
cDNA was used for the synthesis of biotin-labelled cRNA
(Enzo
s
BioArray High Yield RNA Transcript Labeling Kit,
Affymetrix). cRNA purified with RNeasy spin columns from
Qiagen was cleaved into fragments of 35–200 bases by metal-
induced hydrolysis.
Array hybridization and scanning
A measure of 10 mg of biotinylated fragmented cRNA
was hybridized onto the Murine Genome U74Av2 Array
(MG-U74Av2). The array consists of 12 488 probe sets, that
represent RefSeq annotated sequences (B6000) in the Mouse
UniGene database, as well as B6000 EST clones.
The hybridized, washed and coloured arrays were scanned
using the Agilent GeneArray
s
Scanner. Scanned image files
were visually inspected for artifacts and then analysed, each
image being scaled to an all probe set intensity of 150 for
comparison between chips. The Affymetrix
s
Microarray Suite
(version 5.0) was used to control the fluidics station and the
scanner, to capture probe array data and to analyse
hybridization intensity data. Default parameters provided in
the Affymetrix data analysis software package were applied in
running of analyses.
Data analysis
The hybridization values for each gene probe presented on the
array with a set of 16 perfect and mismatch oligonucleotide
pairs were calculated within Affymetrix
s
Microarray Suite 5.0
Software, using the manufacturer’s statistical algorithm. The
results were reported as numeric expression values signal
intensities and absolute information detection calls ‘Present’
or ‘Absent’ produced by two independent algorithms. The
results of a single comparison analysis between two different
arrays were reported for each gene as signal logarithm ratio
(log
2
ratio) and a change call ‘Increase’ or ‘Decrease’. Multiple
data from replicate samples were evaluated and compared
using statistical analyses with the Affymetrix
s
Data Mining
Tool 3.0 (DMT). The average and standard deviation statistics
within Affymetrix
s
DMT was used to summarize the
expression level (the signal values) for each transcript across
the replicates. The unpaired one-sided T-test converting
P-value to a two-sided P-value was used to determine the
direction and significance of change in a transcript’s expression
level between sets of tumours (except for the pool of small
tumours), transgenic liver and normal livers, with the P-value
cutoff determined as 0.05. Besides, only those genes that were
detected (had call ‘Present’) in all samples of a tumour set or
transgenic liver for the upregulated genes and in all control
livers for the downregulated genes were taken into considera-
tion as differentially expressed. Fold-change values were
calculated as the ratio of the average expression levels for
each gene between two tissue sets. Comparison ranking
analysis was additionally employed to study the concordance
of gene expression changes in pairwise comparisons of tumour
samples with control livers. The results are shown as % of
‘Increase’ or ‘Decrease’ calls in individual comparisons, for
example, 16 analyses (four tumours versus four controls) for
Early tumour stage in EGF-induced HCC
J Borlak et al
1818
Oncogene
tumours of median size, 12 (three tumours versus four
controls) for tumours of large size and 12 analyses for
transgenic liver. The small tumours (B1 mm) were pooled
and compared with the four individual controls, resulting in
four comparisons.
Acknowledgements
The excellent technical assistance of Mrs Edith Aretz, Ms Ines
Noack and Mr Albert Rast is gratefully acknowledged. We
thank the Lower Saxony Ministry of Culture and Sciences and
the Volkswagen foundation for providing a grant to J Borlak.
References
Alvarez CV, Shon KJ, Miloso M and Beguinot L. (1995).
J. Biol. Chem., 270, 16271–16276.
Chavrier P, Janssen-Timmen U, Mattei MG, Zerial M, Bravo
R and Charnay P. (1989). Mol. Cell. Biol., 9, 787–797.
Chen S, Zhou B, He F and Yen Y. (2000). Antisense Nucleic
Acid Drug Dev., 10, 111–116.
Choi S, Kobayashi M, Wang J, Habelhah H, Okada F,
Hamada J, Moriuchi T, Totsuka Y and Hosokawa M.
(2000). Clin. Exp. Metastasis, 18, 45–50.
Chung YH, Kim JA, Song BC, Lee GC, Koh MS, Lee YS,
Lee SG and Suh DJ. (2000). Cancer, 89, 977–982.
Clark EA, Golub TR, Lander ES and Hynes RO. (2000).
Nature, 406, 532–535.
Del Sal G, Collavin L, Ruaro ME, Edomi P, Saccone S, Valle
GD and Schneider C. (1994). Proc. Natl. Acad. Sci. USA, 91,
1848–1852.
Del Sal G, Ruaro EM, Utrera R, Cole CN, Levine AJ and
Schneider C. (1995). Mol. Cell. Biol., 15, 7152–7160.
Dey A, Nebert DW and Ozato K. (1991). DNA Cell Biol., 10,
537–544.
Evdokiou A and Cowled PA. (1998). Exp. Cell Res., 240,
359–367.
Gotoh M, Sakamoto M, Kanetaka K, Chuuma M and
Hirohashi S. (2002). Pathol. Int., 52, 19–24.
Hogan B, Beddingtin R, Costantini F and Lacy E. (1994).
Manipulating the Mouse Embryo: A Laboratory Manual.
Cold Spring Harbor Laboratory: Cold Spring Harbor, NY.
Fahmy RG, Dass CR, Sun LQ, Chesterman CN and
Khachigian LM. (2003). Nat. Med., 9, 1026–1032.
Kimura K, Tsuji T, Takada Y, Miki T and Narumiya S.
(2000). J. Biol. Chem., 275, 17233–17236.
Kubicka S, Rudolph KL, Hanke M, Tietze MK, Tillmann
HL, Trautwein C and Manns M. (2000). Liver, 20,
312–318.
Leu JI, Grissey MA and Taub R. (2003). J. Clin. Invest., 111,
129–139.
Melhuish TA, Gallo CM and Wotton D. (2001). J. Biol.
Chem., 276, 32109–32114.
Okabe H, Satoh S, Kato T, Kitahara O, Yanagawa R,
Yamaoka Y, Tsunoda T, Furukawa Y and Nakamura Y.
(2001). Cancer Res., 61, 2129–2137.
Ostrowski J, Woszczynski M, Kowalczyk P, Wocial T, Hennig
E, Trzeciak L, Janik P and Bomsztyk K. (2000). Br. J.
Cancer, 82, 1041–1050.
Ouyang XS, Wang X, Ling MT, Wong HL, Tsao SW and
Wong YC. (2002). Carcinogenesis, 23, 721–725.
Ramljak D, Jones AB, Diwan BA, Perantoni AO, Hochadel
JF and Anderson LM. (1998). Cancer Res., 58, 3590–3597.
Rescan C, Coutant A, Talarmin H, Theret N, Glaise D,
Guguen-Guillouzo C and Baffet G. (2001). Mol. Biol. Cell.,
12, 725–738.
Sandgren EP, Luetteke NC, Qiu TH, Palmiter RD, Brinster
RL and Lee DC. (1993). Mol. Cell. Biol., 13, 320–330.
Schafer DF and Sorrell MF. (1999). Lancet, 35, 1253–1257.
Takeuchi O, Kawai T, Sanjo H, Copeland NG, Gilbert
DJ, Jenkins NA, Takeda K and Akira S. (1999). Gene, 231,
59–65.
Tan JE, Wong SC, Gan SK, Xu S and Lam KP. (2001). J. Biol.
Chem., 276, 20055–20063.
Tatsumoto T, Xie X, Blumenthal R, Okamoto I and Miki T.
(1999). J. Cell. Biol., 147, 921–928.
To
¨
njes RR, Lohler J,
´
Sullivan JF, Kay GF, Schmidt GH,
Dalemans W, Pavirani A and Paul D. (1995). Oncogene, 10,
755–758.
Wang Y, Ripperger J, Fey GH, Samols D, Kordula T, Wetzler
M, Van Etten RA and Baumann H. (1999). Hepatology, 30,
682–697.
Yamaguchi K, Carr BI and Nalesnik MA. (1995). J. Surg.
Oncol., 58, 240–245.
Witz IP. (2000). J. Cell. Biochem. Suppl., 34, 61–66.
Wu JC, Merlino G, Cveklova K, Mosinger Jr B and Fausto N.
(1994). Cancer Res., 54, 5964–5973.
Early tumour stage in EGF-induced HCC
J Borlak et al
1819
Oncogene
... Its expression can be used to distinguish pulmonary lesions originating from primary lung adenocarcinoma or other primaries (Ueno et al., 2003). Additionally, 5 other studies identified IGFBP1, a hepatocytederived secreted protein required for normal liver regeneration by inhibiting proapoptotic signals (Borlak et al., 2005), as an important feature, particularly for the identification of hepatocellular carcinoma, in which it is overexpressed. Wei et al. (2014) identified additional potential cancer signatures not previously associated with the cancer types of interest. ...
Article
Full-text available
Despite recent improvements in cancer diagnostics, 2%-5% of all malignancies are still cancers of unknown primary (CUP), for which the tissue-of-origin (TOO) cannot be determined at the time of presentation. Since the primary site of cancer leads to the choice of optimal treatment, CUP patients pose a significant clinical challenge with limited treatment options. Data produced by large-scale cancer genomics initiatives, which aim to determine the genomic, epigenomic, and transcriptomic characteristics of a large number of individual patients of multiple cancer types, have led to the introduction of various methods that use machine learning to predict the TOO of cancer patients. In this review, we assess the reproducibility, interpretability, and robustness of results obtained by 20 recent studies that utilize different machine learning methods for TOO prediction based on RNA sequencing data, including their reported performance on independent data sets and identification of important features. Our review investigates the strengths and weaknesses of different methods, checks the correspondence of their results, and identifies potential issues with datasets used for model training and testing, assessing their potential usefulness in a clinical setting and suggesting future improvements.
... So bispecific antibody could successfully reduce anti-cancer drug resistance. and EGF-induced liver cancer over-expression in a transgenic mouse model [14,15]. ...
Article
Full-text available
Proto-oncogenes like C-MYC, EGFR and others have physiological function in regeneration, wound and any stressfully injury to maintain tissue Archotexture and healing. Notably, these growth factors work together and had life span to retain to basal level after tissue remodeling and retain its function like what happen in partial hepatectomy. While in cancer, as we work out in 2 transgenic model of liver cancer, we notice that oncogenes do not like each other and just one of them was highly expressive, it keeps other ones at basal level or degraded. Moreover, if one oncogene was inhibited or silenced, other oncogenes become active or expressed and result in anticancer drug resistance. So bispecific antibody could successfully reduce anti-cancer drug resistance.
... Epidermal growth factor (EGF) and transforming growth factor alpha (TGF-alpha) bind to EGFR and trigger MAPK/ERK pathway, thus contributing to MAPK/ERK activation [63]. Multiple lines of evidence suggest induced expression of EGF and TGF-alpha in early-stage hepatocarcinogenesis, demonstrating the possible role of MAPK/ERK pathway in tumor progression of hepatocytes [64,65]. Sorafenib inhibits PI3K/AKT, MAPK signaling cascades leading to mTOR inhibition [58,66]. ...
Article
Full-text available
Hepatocellular carcinoma (HCC) incidence, as well as related mortality, has been steadily increasing in the USA and across the globe, partly due to the lack of effective therapeutic options for advanced HCC. Though sorafenib is considered standard-of-care for advanced HCC, it only improves median survival by a few months when compared to placebo. Sorafenib is also associated with several unpleasant side effects that often lead to early abatement of therapy. Here, we investigate whether a combination regimen including low-dose sorafenib and a non-toxic dose of anti-diabetic drug metformin can achieve effective inhibition of HCC. Indeed, combining metformin with low-dose sorafenib inhibited growth, proliferation, migration, and invasion potential of HCC cells. We observed a 5.3- and 1.9-fold increase in sub-G1 population in the combination treatment compared to sorafenib alone. We found that the combination of metformin enhanced the efficacy of sorafenib and inhibited the MAPK/ERK/Stat3 axis. Our in vivo studies corroborated the in vitro findings, and mice harboring HepG2-derived tumors showed effective tumor reduction upon treatment with low-dose sorafenib and metformin combination. This work sheds light on a therapeutic strategy aiming to augment sorafenib efficacy or dose-de-escalation that may prove beneficial in circumventing sorafenib resistance as well as minimizing related side effects.
... A study showed EGFR was overexpressed in about 68% of human HCC and was correlated with aggressive tumors, metastasis, and poor survival [170][171][172]. A polymorphism in the EGF gene was associated with a high risk for HCC in cirrhotic patients [173,174], and EGF was upregulated in those patients [175,176]. Data on possible communication between EGFR and the cholinergic ligand indicated that overexpression of AChE was detected in HCC cells blocking the activation of MAPK and PI3K/AKT signaling pathway by decreasing the phosphorylation of ERK and AKT; this overexpression also produced an enhancement of the pharmacological effect of druginduced apoptosis [168]. ...
Article
Full-text available
Cancer has been considered the pathology of the century and factors such as the environment may play an important etiological role. The ability of muscarinic agonists to stimulate growth and muscarinic receptor antagonists to inhibit tumor growth has been demonstrated for breast, melanoma, lung, gastric, colon, pancreatic, ovarian, prostate, and brain cancer. This work aimed to study the correlation between epidermal growth factor receptors and cholinergic muscarinic receptors, the survival differences adjusted by the stage clinical factor, and the association between gene expression and immune infiltration level in breast, lung, stomach, colon, liver, prostate, and glioblastoma human cancers. Thus, targeting cholinergic muscarinic receptors appears to be an attractive therapeutic alternative due to the complex signaling pathways involved.
... The stemcell niche environment is rich in other growth factors, such as fibroblast growth factor-2 (FGF-2) and epidermal growth factor (EGF), that induce Id expression through the ERK-MEK pathway in cancer cells by direct binding of early growth responsive protein 1 (Egr-1) to the Id promoter (Borlak et al. 2005;Passiatore et al. 2011;Tournay and Benezra 1996). However, the role of FGF-2 and EGF in Id gene regulation in CNS cells and, in particular, NSPCs is not yet established. ...
Article
Full-text available
Neural stem/progenitor cells (NSPCs) are found in the adult brain and spinal cord, and endogenous or transplanted NSPCs contribute to repair processes and regulate immune responses in the CNS. However, the molecular mechanisms of NSPC survival and integration as well as their fate determination and functionality are still poorly understood. Inhibitor of DNA binding (Id) proteins are increasingly recognized as key determinants of NSPC fate specification. Id proteins act by antagonizing the DNA-binding activity of basic helix-loop-helix (bHLH) transcription factors, and the balance of Id and bHLH proteins determines cell fate decisions in numerous cell types and developmental stages. Id proteins are central in responses to environmental changes, as they occur in CNS injury and disease, and cellular responses in adult NSPCs implicate Id proteins as prime candidates for manipulating stemcell behavior. Here, we outline recent advances in understanding Id protein pleiotropic functions in CNS diseases and propose an integrated view of Id proteins and their promise as potential targets in modifying stemcell behavior to ameliorate CNS disease.
Article
Full-text available
Hepatocellular carcinoma (HCC) is the dominant type of liver cancers and is one of the deadliest health threats globally. The conventional therapeutic options for HCC are hampered by low efficiency and intolerable side effects. Gene therapy, however, now offers hope for the treatment of many disorders previously considered incurable, and gene therapy is beginning to address many of the shortcomings of conventional therapies. Herein, we summarize the involvement of genes in the pathogenesis and prognosis of HCC, with a special focus on dysregulated signaling pathways, genes involved in immune evasion, and non-coding RNAs as novel two-edged players, which collectively offer potential targets for the gene therapy of HCC. Herein, the opportunities and challenges of HCC gene therapy are discussed. These include innovative therapies such as genome editing and cell therapies. Moreover, advanced gene delivery technologies that recruit nanomedicines for use in gene therapy for HCC are highlighted. Finally, suggestions are offered for improved clinical translation and future directions in this area of endeavor.
Article
Full-text available
Hepatocellular carcinoma (HCC) is the most common primary liver cancer. Liver cirrhosis, hepatitis B, hepatitis C, and non-alcoholic fatty liver disease represent major risk factors of HCC. Multiple different treatment options are available, depending on the Barcelona Clinic Liver Cancer (BCLC) algorithm. Systemic treatment is reserved for certain patients in stages B and C, who will not benefit from regional treatment methods. In the last fifteen years, the arsenal of available therapeutics has largely expanded, which improved treatment outcomes. Nevertheless, not all patients respond to these agents and novel combinations and drugs are needed. In this review, we aim to summarize the pathway of trials investigating the safety and efficacy of targeted therapeutics and immunotherapies since the introduction of sorafenib. Furthermore, we discuss the current evidence regarding resistance mechanisms and potential novel targets in the treatment of advanced HCC.
Article
Full-text available
Hepatocellular carcinoma (HCC) is a leading cause of death among cirrhotic patients, for which chemopreventive strategies are lacking. Recently, we developed a simple human cell-based system modeling a clinical prognostic liver signature (PLS) predicting liver disease progression and HCC risk. In a previous study, we applied our cell-based system for drug discovery and identified captopril, an approved angiotensin converting enzyme (ACE) inhibitor, as a candidate compound for HCC chemoprevention. Here, we explored ACE as a therapeutic target for HCC chemoprevention. Captopril reduced liver fibrosis and effectively prevented liver disease progression toward HCC development in a diethylnitrosamine (DEN) rat cirrhosis model and a diet-based rat model for nonalcoholic steatohepatitis-induced (NASH-induced) hepatocarcinogenesis. RNA-Seq analysis of cirrhotic rat liver tissues uncovered that captopril suppressed the expression of pathways mediating fibrogenesis, inflammation, and carcinogenesis, including epidermal growth factor receptor (EGFR) signaling. Mechanistic data in liver disease models uncovered a cross-activation of the EGFR pathway by angiotensin. Corroborating the clinical translatability of the approach, captopril significantly reversed the HCC high-risk status of the PLS in liver tissues of patients with advanced fibrosis. Captopril effectively prevents fibrotic liver disease progression toward HCC development in preclinical models and is a generic and safe candidate drug for HCC chemoprevention.
Article
Full-text available
In normal cells, induction of quiescence is accompanied by the increased expression of growth arrest-specific genes (gas). One of them, gas1, is regulated at the transcriptional level and codes for a membrane-associated protein (Gas1) which is down regulated during the G0-to-S phase transition in serum-stimulated cells. Gas1 is not expressed in growing or transformed cells, and when overexpressed in normal fibroblasts, it blocks the G0-to-S phase transition. Moreover, Gas1 blocks cell proliferation in several transformed cells with the exception of simian virus 40- or adenovirus-transformed cell lines. In this paper, we demonstrate that overexpression of Gas1 blocks cell proliferation in a p53-dependent manner and that the N-terminal domain-dependent transactivating function of p53 is dispensable for Gas1-induced growth arrest. These data therefore indicate that the other intrinsic transactivation-independent functions of p53, possibly related to regulation of apoptosis, should be involved in mediating Gas1-induced growth arrest.
Article
Full-text available
We have analyzed the structure and the regulation of Krox-20, a mouse zinc finger-encoding gene which is transiently activated following serum stimulation of quiescent fibroblast cells in culture. The gene is localized on chromosome 10, band B5, in the mouse, and the homologous human gene also maps to chromosome 10 (region q21.1 to q22.1). Alternative splicing of the 5'-most intron of the Krox-20 gene gives rise to mRNAs encoding putative zinc finger proteins with different N termini. The first exon contains a sequence element with strong similarity to the c-fos proto-oncogene serum response element (SRE). This element can functionally substitute for the c-fos SRE, and it binds the same nuclear protein. It is probably responsible for the serum induction of Krox-20, possibly in combination with a weaker SRE located in the 5'-flanking region of the gene. Our findings suggest that c-fos, Krox-20, and a number of immediate-early serum response genes are coregulated and that the SRE and its cognate protein are essential components of this regulatory pathway.
Article
Full-text available
Transforming growth factor alpha (TGF-alpha) is a polypeptide closely associated with hepatocyte proliferation in vivo and in vitro. In order to investigate the mechanisms by which TGF-alpha contributes to hepatocyte replication and transformation, we isolated hepatocytes from mice bearing a human TGF-alpha transgene and examined their growth properties and gene expression in defined, serum-free culture. The transgenic hepatocytes continued to overexpress human TGF-alpha mRNA and peptide, and were able to proliferate without exogenous growth factors in primary culture, in contrast to nontransgenic mouse hepatocytes. In short-term culture the transgenic hepatocytes underwent 1 wave of DNA replication at 72-96 h in culture before senescing, similar to nontransgenic hepatocytes supplemented with epidermal growth factor. Constitutive expression of TGF-alpha rendered the transgenic hepatocytes unresponsive to further growth stimulation by exogenous TGF-alpha, as well as other mitogens such as epidermal growth factor and hepatocyte growth factor. However, it did not alter their sensitivity to growth inhibition by TGF beta 1, 2 and 3. The addition of nicotinamide to the culture medium enabled both transgenic and epidermal growth factor-supplemented normal hepatocytes to replicate repeatedly and survive for > or = 2 months in primary culture while maintaining differentiated traits. From these long-term primary cultures of transgenic and nontransgenic hepatocytes, we established immortalized cell lines (designated TAMH and NMH lines, respectively). Both lines continued to express differentiated adult hepatocytic markers such as albumin, alpha-1-antitrypsin, transferrin, and connexin 26 and 32 mRNAs, but also expressed mRNAs for the oncofetal markers alpha-fetoprotein and insulin-like growth factor II. Unlike the near-diploid NMH hepatocyte line, the transgenic TAMH hepatocyte line was quasi-tetraploid, strongly expressed human TGF-alpha mRNA, and was highly tumorigenic in nude mice. Well-differentiated hepatocellular carcinomas developed in nude mice given injections of the TAMH line, and these appeared similar to the primary liver tumors seen in TGF-alpha transgenic mice with regard to histology and strong expression of mouse and human TGF-alpha, insulin-like growth factor II, and alpha-fetoprotein mRNAs. Our data show that TGF-alpha overexpression causes autonomous hepatocyte proliferation and contributes to neoplasia but that additional cellular alterations must occur for carcinogenesis. Inappropriate expression of insulin-like growth factor II may constitute one of these steps. The TGF-alpha transgenic mouse hepatocyte line TAMH appears to undergo transformation in a similar manner to that of hepatocytes overexpressing TGF-alpha in vivo, and should serve as an ideal system in which to study hepatocarcinogenesis.
Article
To characterize the effect(s) of transforming growth factor alpha (TGF alpha) during multistage carcinogenesis, we examined tumor development in pancreas and liver of transgenic mice that coexpressed TGF alpha with either viral (simian virus 40 T antigens [TAg]) or cellular (c-myc) oncogenes. In pancreas, TGF alpha itself was not oncogenic, but it nevertheless dramatically accelerated growth of tumors induced by either oncogene alone, thereby reducing the host life span up to 60%. Coexpression of TGF alpha and TAg produced an early synergistic growth response in the entire pancreas together with the more rapid appearance of preneoplastic foci. Coexpression of TGF alpha and c-myc also accelerated tumor growth in situ and produced transplantable acinar cell carcinomas whose rate of growth was TGF alpha dependent. In liver, expression of TGF alpha alone increased the incidence of hepatic cancer in aged mice. However, coexpression of TGF alpha with c-myc or TAg markedly reduced tumor latency and accelerated tumor growth. Significantly, expression of the TGF alpha and myc transgenes in hepatic tumors was induced up to 20-fold relative to expression in surrounding nonneoplastic liver, suggesting that high-level overexpression of these proteins acts as a major stimulus for tumor development. Finally, in both pancreas and liver, combined expression of TGF alpha and c-myc produced tumors with a more malignant (less differentiated) appearance than did expression of c-myc alone, consistent with an influence of TGF alpha upon the morphological character of c-myc-induced tumor progression. These findings demonstrate the importance of TGF alpha expression during multistage carcinogenesis in vivo and point to a major role for this growth factor as a potent stimulator of tumor growth.
Article
Progression of tumor cells toward a high malignancy phenotype and metastasis is a multi‐event cascade involving inter alia alterations in the expression of various genes. The focus of our laboratory is on genes whose altered expression may lead, directly or indirectly, to an increased malignancy phenotype. The identification of such genes and the evaluation of the consequences of their altered expression is essential for attempts to halt tumor progression and prevent metastasis formation. Published work originating in our laboratory showed that members of the murine Ly‐6 supergene family are involved in the progression of certain mouse tumors. The expression level of several members of this family was higher on highly malignant cells than on tumor cells expressing a lower malignancy phenotype. Sorting by flow cytometry of tumor cells to subpopulations expressing either high or low levels of Ly‐6E.1 yielded correspondingly cells expressing a high or a low malignancy phenotype. The high malignancy, high Ly‐6E.1‐expressing cells also expressed high levels of the receptor for urokinase plasminogen activator (uPAR), whereas low malignancy, low Ly‐6E.1‐expressing cells also expressed low levels of uPAR. Transfection studies indicated that uPAR was causally involved in conferring a high malignancy phenotype upon tumor cells expressing high levels of Ly‐6E.1. E48 is a human homologue of the murine ThB protein, a member of the Ly‐6 supergene family (but distinct from the Ly‐6E.1 protein mentioned above) and expressed on head and neck squamous carcinoma cells. Experiments currently in progress are aimed to find out whether E48 is involved in the progression of such cancer cells. Using the differential display technology, it was shown that ligation of E48 on tumor cells by the corresponding antibodies (serving as a surrogate for an as yet unidentified E48 ligand) upregulates an enzyme (FX) involved in the biosynthesis of GDP‐L‐fucose. Fucose is an essential component of certain selectin ligands. J. Cell. Biochem. Suppl. 34:61–66, 2000. © 2000 Wiley‐Liss, Inc.
Article
This study was undertaken in order to assess the main features of hepatocellular carcinoma in Germany, a country with low incidences of this tumor. Two hundred and eighty consecutive patients with hepatocellular carcinomas admitted to the Medical School Hannover between 1993-1997 were retrospectively studied. Reliable data for the assessment of the etiology and the tumor stage of HCC were available for 268 patients. The female/male ratio was 1/4. In 51.9% of the patients, HCC was associated with hepatitis virus B or C (HBV, HCV) infection: 35.1% with HBV, 26.9%) with HCV and 10% coinfection with HBV/HCV This result emphasizes the major impact of HBV and HCV infection in liver cancer in Germany. Of all patients with HCC 74.6%) had liver cirrhosis. The predominant majority of the HCC (87%) were restricted to the liver: in only 5.9% could regional lymph node metastases as well as 8.5%) metastases in other organs be clinically diagnosed by chest X-ray, computed tomography scan or sonography. Data to asses the Okuda tumor stage were available for 166 patients: 47% were classified as stage I, 47% as stage II and only 6% as stage III. Serum AFP were determined in 195 patients. In 66% of the patients, the AFP value was elevated, but only in 30% did the AFP level reach the value of 500 microg/l, which is considered to be significant for HCC diagnosis in patients with liver cirrhosis. The proportion of liver cirrhosis was higher in HCV (97.8%) versus HBV (80.6%) associated HCC, which was the only significant (p<0.05) difference in the characteristics of HCC according to the etiology. Our study shows that liver cirrhosis is the prime risk factor for hepatocarcinogenesis in Germany. However, the very high proportion of hepatitis virus related HCC, in particular the high proportion of HBV infections, contradicts the common view that alcohol is by far the most important etiological factor for hepatocarcinogenesis in low hepatitis virus endemic areas such as Germany.
Article
BACKGROUND Transforming growth factor-α (TGFα) is an important autocrine growth factor of hepatocytes. The authors evaluated the roles of TGFα in chronic viral hepatitis (CVH) and hepatocellular carcinoma (HCC).METHODS The authors measured the amounts of TGFα mRNA in liver tissues from 18 patients with HCC, 31 patients with CVH, and 7 normal controls. “ Hot-start” reverse transcription-polymerase chain reaction (RT-PCR) using oligo-dT and specific primers detected TGFα mRNA in total cellular RNA extracted from liver tissues. The levels of TGFα mRNA were determined by the end point titers of serial, two-fold dilutions of cDNA. The amounts of hepatitis B virus RNA (HBV-RNA) in livers of patients with chronic hepatitis B also were measured by Northern blot hybridization.RESULTSTGFα mRNA levels were extremely higher in patients with HCC compared with patients with CVH and normal controls, and the levels in patients with CVH also were elevated compared with normal controls. The levels of TGFα mRNA were overexpressed in the underlying livers of patients with HCC compared with patients with CVH, although they were lower than those found in HCC tissues. The levels of TGFα mRNA were higher in samples from patients with chronic hepatitis B than in samples from patients with chronic hepatitis C. The levels of TGFα mRNA were not correlated with serum alanine aminotransferase or HBV-RNA levels in liver tissues in patients with chronic hepatitis B. However, the expression of TGFα mRNA tended to be higher in the livers of patients with raised serum α-fetoprotein levels.CONCLUSIONS The overexpression of TGFα mRNA in the liver seems to be associated with the regeneration of hepatocytes rather than hepatic necrosis or viral replication. Also, it may be related closely to the development or progression of HCC, especially in the livers of patients with chronic hepatitis B. Cancer 2000;89:977–82. © 2000 American Cancer Society.
Article
The c-fos proto-oncogene is inducible by cAMP, phorbol esters, serum, and growth factors. The induction by cAMP is mediated by the conserved cAMP response element (CRE), while induction by phorbol esters, serum, and growth factors requires a distal element called the serum response element (SRE). In addition to these elements, a consensus AP-1 transcription factor binding site is located next to SRE. Upstream regions of the mouse and human c-fos genes were footprinted in vivo by the ligation-mediated polymerase chain (PCR). Our results show that all three elements are constitutively protected in mouse liver and lung and in cultured human A431 cells. No major change in the protection profile was detected in A431 cells following stimulation with epidermal growth factor or in mice at birth, when c-fos is known to be induced. These results suggest that the inducible cis elements of the c-fos gene are poised, ready to respond immediately to external signals.