ArticlePDF Available

Transdifferentiation of neoplastic cells

Authors:

Abstract

Transdifferentiation is a process in which a stable cell's phenotype changes to that of a distinctly different cell type. It occurs during certain physiological processes and leads to transition of tumor cell phenotypes. The latter process includes neoplastic epithelial-epithelial transition, neoplastic epithelial-mesenchymal transition, neoplastic mesenchymal-epithelial transition and transition between non-neural and neural neoplastic cell. This phonomenon is exemplified in some origin-debated tumors, such as carcinosarcoma, pleomorphic adenoma, synovial sarcoma, Ewing's/pPNET, and malignant fibrohistiocytoma. We propose that differentiation disturbance of cancer cells should include not only undifferentiation and dedifferentiation, but also transdifferentiation as well. Tumor cell transdifferentiation may be influenced or determined by cellular genetic instabilities, proliferation and apoptosis, as well as by extracellular matrix and growth factors.
Medical Hypotheses (2001) 57(5), 655±666
&2001 Harcourt Publishers Ltd
doi: 10.1054/mehy.2001.1435, available online at http://www.idealibrary.com on
1
Transdifferentiation of
neoplastic cells
Z. Zhang,
1
X.-M. Yuan,
2
L.-H. Li,
1
F.-P. Xie
1
1
Department of Pathology, Dalian Medical University, Dalian, China;
2
Division of Pathology II, Linko
Èping University, Linko
Èping, Sweden
Summary Transdifferentiation is a process in which a stable cell's phenotype changes to that of a distinctly different
cell type. It occurs during certain physiological processes and leads to transition of tumor cell phenotypes. The latter
process includes neoplastic epithelial±epithelial transition, neoplastic epithelial±mesenchymal transition, neoplastic
mesenchymal±epithelial transition and transition between non-neural and neural neoplastic cell. This phonomenon is
exemplified in some origin-debated tumors, such as carcinosarcoma, pleomorphic adenoma, synovial sarcoma, Ewing's/
pPNET, and malignant fibrohistiocytoma. We propose that differentiation disturbance of cancer cells should include not
only undifferentiation and dedifferentiation, but also transdifferentiation as well. Tumor cell transdifferentiation may be
influenced or determined by cellular genetic instabilities, proliferation and apoptosis, as well as by extracellular matrix and
growth factors. &2001 Harcourt Publishers Ltd
INTRODUCTION
Transdifferentiation is a process in which a stable cell's
phenotype changes to that of a distinctly different cell
type. Transdifferentiation may be a kind of differentiation
disturbance occurring in neoplasia, causing the tumor
cells to express a phenotype different from that of their
progenitor (1).
NEOPLASTIC EPITHELIAL±EPITHELIAL
TRANSITION
Normal epithelial transdifferentiation
Transition between different types of epithelium may be
a normal phenomenon in the adult body. For example,
squamous epithelium lining the mouse vagina changes
cyclically in response to sex hormones secreted during the
normal estrous cycle (2).
Epithelial metaplasia
In pathological conditions, mature epithelial cells of one
phenotype change into another, a process called epithelial
metaplasia. Since the resulting cancer cells are immature,
`metaplasia' seems to be a less than accurate term to
describe the transdifferentiation of neoplasia, although
metaplasia is essentially a transdifferentiation process
occurring in non-neoplastic lesions.
Transition of carcinoma cell types
Conversion of carcinoma cell type, resulting in hetero-
genicity of the tumor, is not uncommon. In some cases,
dual-differentiation was found even in individual cells.
The amphicrine carcinomatous cell may exhibit neuro-
endocrine features at its base, and mucus production
and secretory activity in the apical portion (3). Glandular,
squamous, and neuroendocrinal phenotypic features
have been found in the same carcinomatous cell (4).
Chemically induced proliferation of Clara cells in bron-
chioli of the hamster or mouse may initially form
tumors composed of Clara cells, which subsequently
are converted into adeno- or squamous carcinoma
(5,6). Experimental proliferation of hamster bronchial
655
Received 12 January 2001
Accepted 25 May 2001
Corr esponden ce to:P rofessor Zhong Zhang, Pathological Center,
Dalian Medical University, 465 Zhongshan Road, Dalian 116027, P R China.
Fax: 86 0411 472 0610
endothelium and adult bovine or rodent thyroid
follicles can be induced to form mesenchyme when sus-
pended in 3D collagen gels (25). In this process, epithelial
phenotypes such as cytokeratin, cell junction and apical-
based polarity, are lost in the elongated and vimentin-
positive cells, i.e. fibroblasts, which emigrate from the
former surface of the explant. When the cultured cell line
(NBT-II) of a rat bladder carcinoma was attached to the
collagen type I in the medium, or a substitute for serum,
ultroser G, was added to the medium, epithelial-type
cancer cells were converted into the fibroblast type. If the
inducing factors are removed, the epithelial phenotype
may be restored (26).
The effects of differentiation, dedifferentiation and
redifferentiation require a coordinated network that simu-
ltaneously controls cell growth and differentiation. Dif-
ferentiation-induced cells progress through the G
0
-arrest
cycle, whereby a certain population of cells retains the
capacity to de- and redifferentiate and re-enter the cell
cycle. In contrast, the rest of the differentiated population
enters the irreversible G
0
phase (terminal commitment)
that finally results in programmed cell death (27).
Receptor-mediated actions and via intracellular second
messengers cause growth factors to stimulate the cells in
the quiescent G
0
phase of the cell cycle. When the tran-
scription factors are activated, DNA synthesis is initiated
and followed by cell division and dedifferentiation, or
further redifferentiation.
On the other hand, expression of growth factor recep-
tor may influence the regulation of E-cadherin-mediated
cell adhesion. There is increasing experimental evidence
to suggest that the epidermal growth factor receptor of
tyrosine phosphorylation may lead to inactivation of the
E-cadherin±catenin complex in cancer cells, through its
interaction with beta- or gama-catenin in the cytoskeleton
(28). Hepatocyte growth factor-scatter factor (HGF-SF)
has been shown to produce effects similar to those of
epidermal growth factor (EGF) with phosphorylation of
catenins (29). E-cadherin plays a critical role in the
establishment and maintenance of epithelial morphology
and differentiation, and loss of E-cadherin may result in
epithelial±mesenchymal transition.
Cultured neonatel rat hepatocytes (30), mouse mam-
mary gland epithelial cells (NMuMG) (31), nasophanryn-
geal carcinoma cells (CG-I) (32), and rat bladder carcinoma
NBT-II cells (33) undergo epithelial±mesenchymal transi-
tion when the medium contains growth factor, such as
EGF, TGF-alpha, -beta, and some subtypes of fibroblast
growth factor (FGF). Within a few hours of the addition of
nanogram quantities of these factors, the peripheral cells
of NBT-II epithelial islands start to detach and migrate.
Within 10±15 hours, all cells convert into fibroblast-like
cells. This change is reversible by withdrawal of the growth
factor. Although growth factors and their receptors
principally promote cell proliferation and dedifferentia-
tion, some growth factors may also negatively regulate
tumor progression (18).
Transition of carcinoma to carcinosarcoma
Several hypotheses have been proposed to explain the
origin of biphasic appearance of carcinosarcoma. Recent
findings from comparative molecular analysis of the neo-
plastic epithelial and mesenchymal components lend
strong support to the monoclonal multidirectional histo-
genesis of these tumors. Using chromosomal inactivation
assays, k-ras mutation analysis, p53 mutation analysis
and loss of heterozygosity (LOH) studies (34±38), it has
been shown that carcinomatous and sarcomatous
elements in carcinosarcomas share common genetic
alterations, strongly supporting to the monoclonal multi-
directional histogenesis of these tumors. It is generally
thought that there is a single totipotential stem cell, which
gives rise to the multiphasic appearance of carcino-
sarcoma. However, it is highly doubtful whether a specific
totipotential stem cell exists between adult epithelial cells.
Usually both carcinosarcomas and carcinomas occur at
similar sites. If carcinosarcoma originates from totipoten-
tial stem cells, the latter should coexist with the tissue±
determined specific stem cell which gives rise to carci-
noma. Since it is likely that both types of stem cells at the
same location are exposed to the same carcinogen, why is
there such disparity in incidence between carcinoma and
carcinosarcoma?
We know that the primitive mesenchyme derives from
proto-epithelium in the embryo, and that normal and
neoplastic epithelial cells may convert into mesenchyme.
In fact, clinically, carcinoma may progress into carcino-
sarcoma. Some examples of pure endometrial carcinoma-
carcinosarcomatous metastases conversion have been
documented (39). It also has been reported that carcino-
sarcoma may arise in duct papilloma of breast or sweat
gland adenoma (40,41). In two of 16 cases of gynecological
carcinosarcomas with homogeneous or loss of hetero-
geneous (LOH) patterns, Fojii et al. found the specific pat-
terns of genetic progression to be consistent with
sarcomatous components of the neoplasms arising from
carcinomatous components (38). The evidence includes
additional LOH consistent with genetic progression
and diversion from original carcinoma to carcinoma/
sarcoma foci, and additional LOH consistent with genetic
progression from original carcinoma to carcinosarcoma
only in metastatic foci (38). When the cultured rat
hepatocytes were transformed by methyl-nitroso-nitroso-
guanidine and then transplanted into homogenous
rats, they grew into the following tumor types: epidermoid
carcinoma, glandular carcinoma, hepatocarcinoma, undif-
ferentiated carcinoma, sarcoma and carcinosarcoma (42).
Transdifferentiation of neoplastic cells 657
&
2001 Harcourt Publishers Ltd Medical Hypotheses (2001) 57(5), 655±666
NEOPLASTIC MESENCHYMAL±EPITHELIAL
TRANSITION
Mesenchymal±epithelial transition in the embryo
During organogenesis, certain primitive mesenchyme
may regain epithelial phenotype. Metanephric blastema
is composed of highly proliferative mesenchyme.
When the primitive epithelial ureteric bud grows into the
blastema, a reciprocal inductive interaction occurs bet-
ween mesenchyme and epithelium, and the primitive
renal vesicle (epithelial structure) is formed from the
mesenchyme. The formation of renal vesicle reflects
de-repression of epithelial promoters, resulting from
inhibition of the expression or transcriptional activities of
mesenchymal transcription factors.
As mesenchyme undergoes transition to epithelium,
the intermediate filament is changed from vimentin to
cytokeratin, and the non-polarized, loosely associated cells
compact together to form polarized, closely associated
cells. The compaction of cells is mediated by E-cadherin
in a process that is mechanistically analogous to integrin-
mediated spreading of fibroblasts on an extracellular
matrix. E-cadherin expression is crucial for mesenchymal±
epithelial transition. The mechanism controlling
E-cadherin activation probably involves alteration of the
catenin cytoplasmic plaque protein, in the actin cytoske-
leton, or the interactions between them (43,44). If, as
Birchmeir and Behrens (45) pointed out, the endogenous
E-cadherin promoter region is inaccessible in nonepi-
thelial cells, suggesting that the transcriptionally inactiva-
ted gene is packaged into chromatin or is otherwise
modified, e.g. by methylation, then regulation of epithe-
lial-specific expression of E-cadherin largely is due to
specific suppression of promoter activity in nonepithelial
cells, rather than specific activation in epithelial cells.
Expression of epithelial marker in normal and
neoplastic mesenchyme
Decidual stromal cells may express CK8, 18 and 19 when
cultured in either Condimed or Chang conditioned
medium. Similar cytokeratin positivity could also be seen
in cultured fetal fibroblasts from skin, chorionic villi and
lung, but not in young or adult skin fibroblast cultures
using the same cultured conditions (46).
Neoplastic cells co-expressing Vim and CK (and/or
EMA), or expressing E-cadherin, or with epithelial-like
morphology may be found in some primary mesenchymal
tumors.
Scattered cells which were immunocytochemically
positive to CK8 and 18 were found in cultures of Vim-
positive, transformed mesenchymal cell lines, including
SV40-transformed human fibroblasts, rhabdomyosar
coma, rat smooth muscle-derived cells, murine sarcoma
and hamster BHK-2 cells. When cultures of human SV80
fibroblasts were treated with 5-aza-cytidine, the fre-
quency of CK-positive cells increased significantly, but no
cells became positive upon addition of desmosomal
proteins (47).
Using the monoclonal antibody HECD-1, the extra-
cellular domain of E-cadherin immunoreactivity was
identified in a few soft-tissue tumors, especially those with
epithelioid features, including pleomorphic rhabdomyo-
sarcomas, clear cell sarcomas, epithelioid sarcomas,
synovial sarcomas and diffuse mesotheliomas (48).
Treatment with 5-aza-cytidine induces the formation
of a low proportion of CK8-containing intermediate fila-
ments in murine fibroblastoidal or myoblastoidal cells
derived from teratocarcinoma cells, with co-expression of
CK18 and/or 19 in some of those cells. Rare stable cell
lines of these cells display true epithelial morphology with
typical desmosomes and CKs other than 8 and 18 (49).
Effects of tumor suppressor gene and oncogene
The Ela gene of adenovirus may act as a tumor suppressor
gene in certain human tumor cell lines, not by killing
tumor cells, but by converting them into a non-trans-
formed phenotype. Frisch demonstrated that Ela expres-
sion partially convert several human tumor cells
(rhabdomyosarcoma, fibrosarcoma, osteosarcoma) and
fibroblasts into an epithelial phenotype (17).
Co-expression of the human Met receptor (met is an
oncogene) and its ligand, HGF/SF, in NIH3T3 fibroblasts
cause the cells to become tumorigenic in nude mice.
Tsarfary et al. reported that the resultant tumors display
lumen-like morphology, contain carcinoma-like focal areas
with intercellular junctions resembling desmosomes, and
co-express CK and Vim (50).
Whether the oncogene and tumor suppression gene
take part in the mesenchymal±epithelial transition merits
further exploration.
HISTOGENESIS OF SOME
BI-DIFFERENTIATION TUMORS
Pleomorphic adenoma
Currently it is believed that the pleomorphic adenomas
of the salivary gland originate from intercalated duct
epithelium, while myxomatous and chondroid substances
are produced from myoepithelium. Recently, Aigner et al.
(43) found there were some areas with unequivocal epi-
thelial and mesenchymal differentiation in this tumor
tissue. Many areas displayed transitional phenotype cells
and the neoplastic epithelial cells may transdifferentiate
into mesenchymal cells.
658 Zhang et al.
Medical Hypotheses (2001) 57(5), 655±666
&
2001 Harcourt Publishers Ltd
Synovial sarcoma±real carcinosarcoma
Synovial sarcoma rarely originates from the synovial
membrane lining joints and tendon sheaths. The synovial
intima is composed of macrophage-like cells and fibro-
blast-like cells. The flattened, cuboidal or columnar lining
cells of spaces in synovial sarcoma tissue show epithelial
differentiation to an adenomatous, or, in rare instances,
focal squamous, but not synovial intima type. Thus,
synovial sarcoma should be identified as a type of carcino-
sarcoma or carcinoma of mesenchymal tissue (51).
Spindle cell type monophasic, biphasic, and glandular type
monophasic synovial sarcoma may be regarded respec-
tively as sarcoma, carcinosarcoma and carcinoma in the
differentiation spectrum of this neoplasm.
Epithelioid sarcoma
In view of the immunohistochemical evidence, epithelioid
sarcoma may be hypothesized to be a type of mesenchy-
mal carcinoma with simple epithelial differentiation.
Electron microscopic photos of the tumor showed a
spectrum of cellular differentiation from fibrohistiocytic
cells to epithelial-type cells with junctions, microvilli and
tonofilaments. Spindle cells showed myofibroblastic and
fibroblastic differentiation (52). All the tumors display
both vimentin and epithelial markers, just like some
poorly differentiated carcinoma cells. In some cases the
tumor cells may co-express neurofilament protein,
neuron-specific enolase (NSE), or even synaptophysin (53).
Karyotypes of the RM-HS1 cell line were reported to
contain extensive numerical and structural rearrange-
ments, with up to 24 marker chromosomes (54). Quezado
reported allellic loss on chromosome 22q in six of 10
epithelioid sarcomas (55).
Malignant mesothelioma
The regenerating mesothelium may originate from
either the surrounding uninjured mesothelial cell popu-
lation or from subserosal cells. It has been suggested
that subserosal cells are distinct from other connective
tissue cells in this ability to co-express Vim and CK, and
serve as replicated cells which can differentiate into
surface epithelium. To study this phenomenon, both the
visceral and parietal peritonea of rats was incised and
allowed to heal with resultant proliferation and differ-
entiation of subserosal cells. The proliferated subserosal
cells change morphologically from the original spindle-
shaped, fibroblast-like cells to polygonal surface epithelial
cells (56). Following injection of crocidolite asbestos
into the peritoneal cavity of rats and mice, the subserosal
cells proliferated to form a tumorous nodule, in which
the spindle cells differentiated into surface mesothelial
cells. No evidence of proliferation by the original surface
mesothelial cells in response to the asbestos was found
(56). Most malignant mesotheliomas are characterized
by multiple, complex and heterogeneous chromosomal
aberrations. Sreekantaiah et al. reported that the 3p
abnormalities, perhaps causally related to the develop-
ment of this malignancy, and del(6q) might represent
a primary change (57). Malignant mesothelioma originate
from subserosal mesenchymal cells. As its normal pro-
genitor possesses bidifferentiated potential, this tumor
may show a differentiation spectrum from sarcoma to
carcinosarcoma and carcinoma.
Adamantinoma of long bone
Adamantinoma of long bone, composed of epithelial and
fibrous components, is closely related to osteofibrous
dysplasia (OFD). Most OFDs contain isolated or aggregated
keratin-positive cells, which are identical to OFD-like
adamantinoma. The recurrent focus of OFD-like tumor
with isolated keratin-positive cells may be a classic ada-
mantinoma with abundant epithelium (58). Hazelbag
et al. found: (a) focal basement membrane substances in
OFD-like areas with keratin-positive cells; (b) increased
basement membrane continuity with gain of histological
distinction between epithelial and fibrous components;
and (c) strong tenascin reactivity directly surrounding
well-developed epithelial fields, with weaker staining
more distantly (59). Hazelbag et al. suggested that the
epithelial component in adamantinoma may be trans-
formed directly from osteofibrous tissue (59).
HISTOGENESIS OF EWING'S/pPNET
Hypothesis on the origin of Ewing's sarcoma
Ewing's/pPNET (Ewing's sarcoma/peripheral primitive
neuroectodermal tumor) is a family of bone and soft-
tissue tumors, in which the typical Ewing's sarcoma, with
a mesenchymal phenotype, lies at one end of the differ-
entiation spectrum and pPNET, with clear evidence of
a neural marker, lies at the other end. Some authors had
suggested that Ewing's sarcoma originates from the
mesenchymal or primitive form of connective tissue (60).
Aurias et al. and Wang-Peng et al. found a coincidental
cytogenetic abnormality in ES and pPNET (60). In 1990
several groups of investigators found that, through the
addition of agents such as retinoic acid to the cultures, the
ES cells may be induced in vitro to express neural markers
(60). Currently, many pathologists prefer to support the
view that Ewing's/pPNET originates from neural crest
cells. The embryonic neural crest has the potential to
differentiate into mesenchyme, which could explain the
expression of Ewings' mesenchymal phenotype.
Transdifferentiation of neoplastic cells 659
&
2001 Harcourt Publishers Ltd Medical Hypotheses (2001) 57(5), 655±666
Against the hypothesis of a neural origin of
Ewing's/pPNET
Although normal mesenchyme does not express neural
differentiation, `neural' markers may be found in mesen-
chymal tumors. For example, epithelioid sarcoma cells
express neurofilament, NSE and synaptophysin protein
(61); MFHs have neurofilaments and are neural-associated
antigen-positive (62); synovial sarcoma may be immuno-
histochemically positive to neurofilament 68kd, S-100,
Leu-7 and GFAP (63).
Molenaar and Muntinghe (64) studied the expression of
neural cell adhesion molecules and neurofilament protein
isoforms in bone and soft-tissue sarcomas. They found
that both neural markers are expressed in rhabdomyo-
sarcoma, leiomyosarcoma, fibrosarcoma, MFH, malignant
rhabdoid tumor and fibromatosis, and are N-CAM-
positive in synovial sarcoma. Desmoplastic small round
cell tumors (DSRCT) may show positive staining with
antibodies to NES, Leu-7 and other neural antigens,
although they are more likely from mesenchymal tissue
such as subserosal cells (65). Thus, the expression of
`neural' markers does not prove a neural origin for the
tumor.
Neuroblastoma originates from the sympatho-adrenal
lineage of the neural crest. Its characteristic chromosome
changes are deletion of 1p36.2±3, amplification of the
proto-oncogene MYCN, and abnormalities of the chromo-
some number (66). The special cytogenetic aberration
found in Ewing's/pPNET is chromosomal translocation
resulting in gene fusion. One of the criteria used to diag-
nose classic neuroblastoma is the presence of Homes±
Wright pseudorosettes. However, the number of rosettes
seen by most investigators in pPNETs is few or none (60).
The differentiation-potential of neuroblastoma shows
neural mono-direction. Ewing's/pPNET shows multi-
differentiation and may overlap somewhat in appearance
with DSRCT, or even alveolar rhabdomyosarcoma, both
of which also show characteristic chromosome translo-
cation. From these points, Ewing's/pPNET is more closely
related to non-neural derived tumor in phenotype and
cytogenetic profile. The discovery of a hybrid tumor
comprised of Ewing's/pPNET and DSRCT further demon-
strated that DSRCT and pPNET may share a common
histogenesis (67).
Neural crest cells detach from the dorsal neural tube
and migrate, colonize and differentiate into a large
variety of derivatives. These include the neurons and glial
cells of the peripheral nervous system, endocrine cells,
melanocytes and the so-called mesoectoderm, which
plays a crucial role in formation of the connective tissues
and skeletal structures of the head. Most of the Ewing's/
pPNET were located outside the head, neck and regions
unrelated to any specific structures of the peripheral or
sympathetic nervous systems. These tumors are thought
to originate from abnormally migrated neural crest cells.
However, the occurrence of Ewing's/pPNET in adult and
aged patients suggests that an implanted neuroecto-
dermal fetal origin of these tumors is unlikely. Fukunaga
et al. reported a case of carcinosarcoma with neuroecto-
dermal differentiation which occurred in the uterus of
a 54-year-old woman (68). The neoplastic elements
included squamous carcinoma, leiomyosarcoma and
islands of small- to medium-sized cells with rosette-like
formation and immunoreactivity for GFAP, Leu-7, NSE
and synaptophysin. Certainly in this case, the cells with
neural differentiation did not arise from neural crest cell.
Relation between chromosome translocation and
morphogenesis in Ewing's/pPNET
Ewing's/pPNET is characterized by specific chromosomal
translocations, which result in a fusion gene and its
encoding chimeric protein. In at least 90% of Ewing's/
pPNET, the translocation is t(11;22) (q24;q12), resulting in
fusion of parts of the EWS gene with parts of the FLI-1 gene
encoding for transcription factor. In the encoded chimeric
protein, the amino-terminal EWS domain is linked to the
DNA-binding domain of the transcription factor. In
another 5% of cases, t(21;22) (q22;q12) translocation is
present, which involves the genes for EWS and ERG.
Another rare translocation is t(7;22) (p22;q12), which
results in EWS/ETV-1. FLI-1, ERG and ETV-1 all belong
to the ETS proto-oncogene family (69).
EWS/FLI-1 (or EWS/ERG, ETV-1) is a transcription
factor. May et al. demonstrated that EWS/FLI-1 efficiently
transformed NIH3T3 cells, but FLI-1 did not (70).
Experiments with EWS/FLI-1 deletion mutants indicated
that both EWS and FLI-1 domains are necessary for
transformation by the t(11;22) translocation product
(71). Recently, Teitell et al. reported that NIH3T3 fibro-
blasts infected with either EWS/FLI-1 or EWS/ETV-1
resembled small round cell tumor microscopically, and
showed both epithelial and neuroectodermal pheno-
types as represented by CK15 expression, dense junc-
tion, and neurosecretory granule formation (72). The
altered cells lost both extracellular collagen deposi-
tion and RER, indicating a loss of mesenchymal
differentiation.
Although all the Ewing's/pPNETs tumors harbor an
identical model of the chromosomal translocation, its
differentiated spectrum includes a mesenchymal pheno-
type at one end, and a neural at the other. Other small cell
tumors of soft tissue, such as DSRCT, polyphenotypic
tumors, and even a few cases of alveolar or embryonal
rhabdomyosarcoma, may also possess t(11;22) (q24;q12)
translocation and the EWS/FLI-1 fusion gene (73,74). From
these findings, we deduce that expression of the neural
660 Zhang et al.
Medical Hypotheses (2001) 57(5), 655±666
&
2001 Harcourt Publishers Ltd
phenotype in Ewing's/pPNETs is not decided merely by
the formation of EWS/FLI-1 or EWS/ERG fusion gene, but
also requires the activation of certain genes related to
neural differentiation.
Difference in origin between cPNET and
Ewing's/pPNET
There are two categories of primitive neuroectodermal
tumors (PNETs): central (cPNET ) and peripheral (pPNET)
categories (75). A tumor which originates in the brain or
spinal cord, such as medulloblastoma, is classified as a
central PNET. Adrenal and extra-adrenal neuroblastomas
and Ewing's/pPNET (in soft tissue and bones) are exam-
ples of a peripheral PNET. Because they occur in these
specific locations, central PNETs and neuroblastomas may
derive from neural tissue. As for Ewing's/pPNET, we
hypothesize that they arise from mesenchymal tissue. In
view of the preceding part of this review, the reasons may
be summarized as follows: expression of neural markers
does not signify a neural origin; the primary location of
most tumors is unrelated to any specific structure of the
nervous system; its phenotype is distinct from that of
neuroblastoma but similar to that of some small cell tumors
of the mesenchyme; its characteristic chromosome trans-
location and fusion gene, with which the infected 3T3
fibroblasts gain neuroectodermal phenotypes and loss of
mesenchymal differentiation. It is evident that the func-
tion of EWS/FLI-1 or EWS/ERG and EWS/ETV-1 is not only
to initiate the transformation of mesenchymal cells, but
also to regulate the phenotypic expression of Ewing's/
pPNET.
Vasoactive intestinal peptide (VIP) is a neuromodulator
which regulates both proliferation and differentiation of
neuronal precursors. Fruhwald et al. analyzed cPNET cell
lines, cPNET tumors and Ewing's/pPNET, using reverse
transcriptase-polymerase chain reaction and Southern
hybridization (76). They found that VIP receptor 1 (VIPR1)
and VIPR2 are more highly expressed in both primary
cPNET tumors and cPNET cell lines. They pointed out
that this result may reflect the divergent pathways of
cPNET toward neural differentiation and Ewing's/pPNET
to mesenchymal cells. The question is how such divergent
pathways of differentiation occur. Another possible
explanation is that the origin of cPNET is distinct from
that of pPNET; the former from neural tissue, and the
latter from mesenchymal cells.
HISTOGENESIS OF MALIGNANT
FIBROHISTIOCYTOMA
Malignant fibrohistiocytoma (MFH) is composed
of fibroblast-like cells, histiocyte-like cells and an
intermediate cell type (77,78). These neoplastic cells may
show myofibroblast-like differentiation, occasionally
acquire certain properties of osteoblasts or chondroblasts,
and immunohistochemically express CK, epithelial
membrane antigen (EMA) neurofilaments (NF) and
neural-associated antigens (79±82).
Heterogeneity of MFH
Histocyte origin
Since the cases studied did not react with anti-monocyte-
macrophage antibodies, some investigators are moving
away from histiocyte-origin hypothesis (83±88). However,
other investigators have reported that MFH does express
antigenic markers specific for monocytes-macrophages
(89±91). Yumoto and Morimoto inoculated SV40-trans-
formed bone marrow macrophages in syngeneic mice to
produce a transplantable tumor (92). The tumor was
composed of spindle cells with histiocytic function such
as storiform arrangement. Some tumor cells present a
transitional form between histiocyte-like and fibroblast-
like cells. This experiment demonstrated that MFH may be
derived through transdifferentiation of histiocytes.
Fibroblast origin
This theory is supported by substantial evidence: MFH
cells expressed mesenchymal antigens (84,85,88). Injec-
tion of DMBA into the knee-joint cavity of rats may induce
MFHs and fibrosarcomas or synovial sarcomas in and
around the joint (93,94). After several recurrences, fibro-
sarcoma may transform into MFH (95). Fibrosarcoma cells
(RFS ) derived from a cadmium-induced fibrosarcoma of
rat, and RFS cells in different subcultures produced
MFH and fibrosarcoma in nude mice and baby rats
(95). Non-transformed and transformed fibroblast-like
mouse embryo cells can be induced to differentiate into
macrophages or histiocytes in a medium supplemented
with human plasma (96). These findings demonstrated
that histiocyte-like cells could result from transdiffer-
entiation of neoplastic fibroblasts.
Other origins
The dedifferentiated components of dedifferentiated sar-
coma (e.g. dedifferentiated liposarcoma, dedifferentiated
chondrosarcoma and dedifferentiated leiomyosarcoma)
and sarcomatous cells of carcinosarcoma, e.g. sarcoma-
tous carcinoma of lung, kidney and prostate and carci-
nosarcoma of breasts, usually look like MFH or
fibrosarcoma (97±105). In addition, the individual case of
malignant melanoma, gliosarcoma and malignant lym-
phoma, has the same appearance as MFH on microscopic
or even electron-microscopic examination (106±108).
Transdifferentiation of neoplastic cells 661
&
2001 Harcourt Publishers Ltd Medical Hypotheses (2001) 57(5), 655±666
Karyotype progression of MFH
The karyotype of MFH is far more complex than that of
fibrosarcoma. Cytogenetic evidence of clonal evolution
has been found in some tumors (109). More than half the
tumors were near triploid or near tetraploid, some were
hyperdiploid and even hypodiploid (110). The tumors
develop through acquisition of structural aberrations and
chromosome loss or through gain of a chromosome (111).
Ring, dicentric- and telomeric-associated chromosomes
seem to represent early events in the development of
MFH (109,112,113). Sometimes, the presence of ring
chromosomes is the sole cytogenetic aberration of myxoid
MFH (112,113), as well as the sole consistent chro-
mosomal alteration of well-differentiated liposarcoma
and dedifferentiated liposarcoma with area of MFH ele-
ments (114±116). Ring, dicentric- and telomeric-associ-
ated chromosomes result from telomeric erosion and
are highly unstable, leading to further breakage and
fusion (117).
Genberg et al. established two cell lines from two sub-
sequent recurrences of MHF with identical marker chro-
mosome in common. The U-2149 cell line from the second
recurrence consisted of mainly fibroblast-like cells. The U-
2197 cell line from the third recurrence was composed of
mainly histiocyte-like cells. Based on their data, Genberg
et al. explained the appearance of histiocyte-like cells in
MFH as a consequence of chromosomal progression (118).
APOPTOSIS AND TRANSDIFFERENTIATION
Apoptosis is a universal cellular suicide pathway in
normal elimination and development. Disruption of this
normal process resulting in illegitimate cell survival can
facilitate cancer development and progression.
In histological tumor material apoptotic index (AI) may
be used as a measure of the extent of apoptosis. Most often
AI is defined as a percentage of apoptotic cells and bodies
per total number of tumor cells. A consistent feature in
many studies is the positive correlation or association
between apoptosis and proliferation. In general, rapidly
growing tumors have a greater degree of apoptosis than
relatively indolent ones. Some investigators have begun
investigating the correlation between apoptosis and neo-
plastic transdifferentiation.
The Bcl-2 family and apoptosis in synovial sarcoma
The very recent report of Antonescu et al. (119) indicated
that the monophasic types of synovial sarcoma (with SYT-
SSX2) have a significantly higher mean and median ki-67
labeling index than biphasic types (with SYT-SSX1) (119),
and apoptosis as assessed by TUNEL was rarely observed
in both types, consistent with prominent expression of the
anti-apoptosis protein Bcl-2 in almost all cases. Bcl-2 pro-
tein expression was strong in spindle cells in contrast to
the weak or negative reactivity observed in epithelial cells.
Bcl-2-positive spindle cells consistently surrounded the
negative epithelial elements, a pattern confirmed by the
studies of Antonescu et al. and other investigators (119).
The spindle and adenomatous elements in biphasic syno-
vial sarcoma may be analogous to the undifferentiated
and well-differentiated elements of carcinoma, confirming
that bcl-2 protein expression is indeed different between
two areas of the biphasic tumor.
p53 mutation or inactivation, and apoptosis in MFH
The p53 tumor-suppressor gene is the most commonly
mutated gene in human cancers. Two biological functions
are attributed to p53: regulation of cell cycle progression
after stress, and regulation of apoptosis.
Some evidence indicates that, although p53 mutations
are unlikely to be the primary cause, they may exacerbate
chromosome instability (120). Pilotti et al. investigated
p53 and MDM2 over-expression in 98 lipomatous tumors
by immunocytochemistry, as well as by molecular and
cytogenetic analysis (121). Among the 74 cases of lipo-
sarcomas, 14 cases were of the dedifferentiated subtype
in which the differentiated component was well
differentiated, while the dedifferentiated area had MFH
features in 13 cases. The results show that, in the retro-
peritoneal well-differentiated±dedifferentiated group,
MDM2-mediated inactivation of p53 could be related to
the mechanism of transdifferentiation, while, in the non-
retroperitoneal WD-DD group, the Tp53 mutations appear
to correlate with the dedifferentiation process. Hisaoka
et al. studied eight cases of dedifferentiated liposarcoma
with myxoid MFH areas (122). The clinicopathologic,
cytogenetic and molecular features showed that the dif-
ferentiated portion of tumors was more closely related
to well-differentiated liposarcoma rather than to ordinary
myxoid liposarcoma, while the dedifferentiated part
was myxoid MFH. The myxoid portions frequently
showed a higher proliferative activity, and more MDM-2
and p53-positive tumor cells than in well-differentiated
areas. Hisaoka et al. claimed that an altered p53 pathway,
including p53 gene mutations and MDM-2-mediated
inactivation of p53, may play a role in tumorigenesis of
this myxoid subtype of liposarcoma and its progression
(dedifferentiation) toward conversion into myxoid
MFH (122).
CONCLUSION
Differentiation disturbance of cancer cells should include
not only undifferentiation and dedifferentiation, but also
662 Zhang et al.
Medical Hypotheses (2001) 57(5), 655±666
&
2001 Harcourt Publishers Ltd
transdifferentiation. A spectrum of phenotypic features
has been identified in various embryonal carcinoma cells,
which resemble normal embryonic cells at different stages
of development (123). Just like embryonal carcinoma,
most cancers are similar in phenotype to their normal
forebears, but lower in differentiation, e.g. dedifferenti-
ation or even undifferentiation. On the other hand, in
some tumors, as compared to their normal progenitor
cells, the morphophenotypes of tumor cells are converted
into other types by so-called transdifferentiation.
The pathogenesis and progression of neoplasm are
closely associated with genetic instabilities, including sub-
tle sequence changes, alteration in chromosome number,
chromosome translocation and gene amplification (120).
Genetic instabilities are the basis of neoplastic dediffer-
entiation and transdifferentiation. Neoplastic epithelial±
epithelial transition or neoplastic epithelial±mesenchymal
transition may be simply the result of subtle sequence
changes and alteration in extracellular matrix molecules
or stimulation by growth factors, via regulation of cell
proliferation, differentiation and programmed cell death.
As to neoplastic mesenchymal±epithelial transition in
synovial sarcoma and mesenchymal±neural transdiffer-
entiation in Ewing's/pPNET, chromosome translocations
(gene rearrangements) are crucial for transdifferentiation.
The histological origin of some groups of neoplasms
has been debated for decades. Debated tumors include:
multidirectional differentiation tumors; tumors with the
characteristic phenotype of a certain tissue occurring in
an organ foreign to such tissue, e.g. meningioma in lung;
and tumors with a phenotype dissimilar to any type of
normal tissue, e.g. alveolar soft part sarcoma. The histo-
logical appearances of these tumors may result from
transdifferentiation of certain neoplastic cells, and their
histogenesis may be resolved at the cytogenetical and
molecular levels.
REFERENCES
1. Zhang Z., Xie F.-P. Transdifferentiation of neoplastic cell. Foreign
Medical Science (Oncology) (in Chinese) 1994; 21 (suppl): 8±9.
2. Horvat B., Vrcic H., Damjanov I. Transdifferentiation of murine
squamous vaginal epithelium in proestrus is associated with
changes in the expression of keratin polypeptides. Exp Cell Res
1992; 199: 234±239.
3. Hammer S. P., Insalaco S. J., Lee R. B. Amphicrine carcinoma of
the uterine cervix. Am J Clin Pathol 1992; 97: 516±522.
4. McDowell E. M., Trump B. F. Pulmonary small cell carcinoma
showing tripartite differentiation in individual cells. Hum Pathol
1981; 12: 286.
5. Rehm S., Takahashi M., Ward J. M. et al. Immunohistochemical
demonstration of Clara cell antigen in lung tumors of bronchiolar
origin induced by N-nitrosodiethylamine in Syrian golden
hamsters. Am J Pathol 1989; 134: 79±87.
6. Rehm S., Lijingsky W., Singh G. et al. Mouse bronchiolar
cell carcinogenesis. Histologic characterization and expression
of Clara cell antigen in lesions induced by N-nitrosobin-
(2-chloroethyl) ureas. Am J Pathol 1991; 139: 413±422.
7. Gould V. E. Histogenesis and differentiation. A re-evaluation of
these concepts as criteria for the classification of tumors.
Hum Pathol 1986; 17: 212±215.
8. Xie Z. X., Liao J. H., Zheng P. E. Immunohistochemical and
electromicroscopic observation for endocrine and Paneth's cell of
gastric carcinoma. Chin J Pathol (in Chinese) 1992; 21: 221±223.
9. Wu W. Y., Sun Y. N., Liu Q. et al. Immunohistochemical study
on bladder transitional cell carcinoma containing HCG. Chin J
Pathol (in Chinese), 1989; 18: 34±36.
10. Jiang S. X., Yang G. H., Tan Y. S. et al. Study on cellular
heterogeneity of bronchiogenic carcinoma. Chin J Pathol
(in Chinese) 1991; 20: 178±180.
11. Iemura A., Yano H., Mizogushi A. et al. A cholangiocellular
carcinoma nude mouse strain showing histologic alteration from
adenocarcinoma to squamous cell carcinoma. Cancer 1992; 70:
415±422.
12. Cox W. F., Pierce G. B. The endodermal origin of the endocrine
cells of an adenocarcinoma of the colon of the rat. Cancer 1982;
50: 1530±1538.
13. Kirkland S. C. Clonal origin of columnar, mucous, and endocrine
cell lineages in human colorectal epithelium. Cancer 1988; 61:
1359±1363.
14. de Bruine A. P., Dinjens W. N. M., van den Linden E. P. M. et al.
Extracellular matrix components induce endocrine
differentiation in vitro in NCL-H716 cells. Am J Pathol 1993; 142:
773±782.
15. Fitchet J. E., Hay E. D. Medial edge epithelium transforms to
mesenchyme after embryonic palatal shelves fuse. Dev Biol 1989;
131: 455±474.
16. Griffith C. M., Hay E. D. Epithelial-mesenchymal transformation
during palatal fusion: Carboxyfluorescein traces cells at light and
electron microscopic levels. Development 1992; 116: 1087±1099.
17. Frisch S. M. Ela induces the expression of epithelial
characteristics. J Cell Biol 1994; 127: 1085±1096.
18. Thiery J. P., Chopin D. Epithelial cell plasticity in development
and tumor progression. Cancer Metast Rev 1999; 18: 31±42.
19. Ward J. M., Stevens J. L., Konishi N. et al. Vimentin metaplasia in
renal cortical tubules of preneoplastic, neoplastic, aging and
regenerative lesions of rats and humans. Am J Pathol 1992; 141:
955±964.
20. Gould V. E., Koukoulis G. K., Jansson D. S. et al. Coexpression
patterns of vimentin and glial filament protein with cytokeratins
in the normal, hyperplastic and neoplastic breast. Am J Pathol
1990; 137: 1143±1156.
21. Domagala W., Lasota J., Bartkowiad J. et al. Vimentin is
preferentially expressed in human breast carcinomas with
low-estrogen receptor and high Ki-67 growth fraction. Am J
Pathol 1990; 136: 219±227.
22. Eguchi G., Okada T. S. Differentiation of lens tissue from the
progeny of chick retinal pigment cells cultured in vitro: A
demonstration of a switch of cell type in clonal cell culture.
Proc Natl Acad Sci USA 1973; 70: 1495±1499.
23. Agata K., Kobayashi H., Itoh Y. et al. Genetic characterization of
the multipotent dedifferentiated state of pigmented epithelial
cells in vitro. Development 1993; 118: 1025±1030.
24. Eguchi G., Kodama R. Transdifferentiation. Curr Opin Cell Biol
1993; 5: 1023±1028.
25. Hay E. Extracellular matrix alters epithelial differentiation.
Curr Opin Cell Biol 1993; 5: 1029±1035.
26. Boyer B., Tucker G. C., Valles A. M. et al. Reversible transition
towards a fibroblastic phenotype in a rat carcinoma cell line.
Int J Cancer 1989; (Supple) 4: 69±75.
Transdifferentiation of neoplastic cells 663
&
2001 Harcourt Publishers Ltd Medical Hypotheses (2001) 57(5), 655±666
27. Hass R. Retrodifferentiation and cell death. Criti Rev Onco-genesis
1994; 5: 359±371.
28. Jawhari A. U., Farthing M. J. G., Pignatelli M. The E-cadherin/
epidermal growth factor receptor interaction ± A hypothesis of
reciprocal and reversible control of intercellular adhesion and cell
proliferation. J Pathol 1999; 187: 155±157.
29. Crepaldi T., Pollack A. L., Part M. et al. Targeting of the SF/HGF
receptor to the basolateral domain of polarized epithelial cells.
J Cell Biol 1994; 125: 313±320.
30. Pagan R., Martin I., Alonso A. et al. Vimentin filaments follow
the pre-existing cytokeratin network during epithelial
mesenchymal transition of cultured neonatal rat hepatocytes.
Exp Cell Res 1996; 222: 333±334.
31. Miettnen P. J., Ebner R., Lopez A. R. et al. TGF-beta induced
transdifferentiation of mammary epithelial cells to mesenchymal
cells: involvement of type I receptors. J Cell Biol 1994; 127:
2021±2036.
32. Chen J. K., Shen L. S., Chao H. H. Correlation of transformation
from epithelial to mesenchymal-like morphology and
endogenous bFGF levels in human nasopharyngeal carcinoma
cells. J Cellular Physiol 1994; 160: 401±408.
33. Valle A. M., Boyer B., Badet J. et al. Acidic fibroblast growth factor
is a modulator of epithelial plasticity in a rat bladder carcinoma
cell line. Proc Natl Acad Sci USA 1990; 87: 1124±1128.
34. Thompson L., Chang B., Barsky S. H. Monoclonal origins of
malignant mixed tumors (carcinosarcomas). Evidence for a
divergent histogenesis. Am J Surg Pathol 1996; 20: 277±285.
35. Wada H., Enomoto T., Fujita M. et al. Molecular evidence that
most but not all carcinosarcomas of the uterus are combination
tumors. Cancer Res 1997; 57: 5379±5385.
36. Kounelis S., Jones M. W., Papadaki H. et al. Carcinosarcomas
(malignant mixed mullerian tumors) of the female genital tract:
Comparative molecular analysis of epithelial and masenchymal
components. Hum Pathol 1998; 29: 82±87.
37. Abeln E. C., Smit V. T. H. B. M. Wessels J. W. et al. Molecular genetic
evidence for the conversion hypothesis of the origin of malignant
mixed mullerian tumours. J Pathology 1997; 183: 424±431.
38. Fujii H., Yoshida M., Gong Z. X. et al. Frequent genetic
heterogeneity in the clonal evolution of gynecological
carcinosarcoma and its influence on phenotypic diversity.
Cancer Res 2000; 60: 114±120.
39. Sreenan J. J., Hartn W. R. Carcinosarcomas of the female genital
tract. A pathologic study of 29 metastatic tumors: Further
evidence for the dominent role of the epithelial component and
the conversion theory of histogenesis. Am J Surg Pathol 1995;
19: 666±670.
40. Pitt M. A., Swells S., Eyden B. P. Carcinosarcoma arising in a duct
papilloma. Histopathol 1995; 26: 81±84.
41. Mckee P. H., Fletcher C. O. M., Stavrino P. et al. Carcinosarcoma
arising in eccrine spiradenoma (a clinicopathologic and
immunohistochemical study of 2 cases). Am J Dermatopathol
1990; 12: 335±343.
42. Tsao M. S., Grisham J. W. Hepatocarcinomas,
cholangiocarcinomas, and hepatoblastomas produced by
chemically transformed cultured rat liver epithelial cells. Am J
Pathol 1987; 127: 169±181.
43. Aigner T., Neureiter D., Volker U. et al. Epithelial-mesenchymal
transdifferentiation and extracellular matrix gene expression in
pleomorphic adenomas of the parotid salivary gland. J Pathol
1998; 186: 178±185.
44. Gumbiner B. M. Cell adhesion: The molecular basis of tissue
architecture and morphogenesis. Cell 1996; 84: 345±357.
45. Birchmeier W., Behrens J. Cadherin expression in
carcinoma: Role in the formation of cell junction and the
prevention of invasiveness. Biochem Biophys Acta 1994; 198:
11±26.
46. von Koskull H., Virtanen I. Induction of cytokeratin
expression in human mesenchymal cells. J Cell Physio 1987;
133: 321±329.
47. Knapp A. C., Franke W. W. Spontaneous losses of control of
cytokeratin gene expression in transformed, non-epithelial
human cells occurring at different levels of regulation. Cell 1989;
59: 67±79.
48. Sato H., Hasegawa I., Abe Y. et al. Expression of E-cadherin in bone
and soft tissue sarcomas: A possible role in epithelial
differentiation. Hum Pathol 1999; 30: 1344±1349.
49. Semat A., Duprey P., Vasseur M. et al. Mesenchymal-epithelial
conversions induced by 5-azacytidine: appearance of cytokeratin
endo-A messenger RNA. Differentiation 1986; 31: 61±66.
50. Tsarfary I., Rong S., Ressau J. H. et al. The Met proto-oncogene
mesenchymal to epithelial cell conversion. Science 1994; 263:
98±101.
51. Chadially F. N. Is synovial sarcoma a carcinosarcoma of
connective tissue? Ultrastruct Pathol 1987; 11: 147±151.
52. Fisher C. Epithelioid sarcoma. The spectrum of ultrastructural
differentiation in seven immunohistochemically defined cases.
Hum Pathol 1988; 19: 265±275.
53. Gerharz C. D., Moll R., Ramp U. et al. Multidirectional
differentiation in a newly established human epithelioid sarcoma
cell line (GRU-1) with co-expression of vimentin, cytokeratins and
neurofilament proteins. Int J Cancer 1990; 45: 143±152.
54. Reeves B. R., Fisher C., Smith S. et al. Ultrastructural,
immunocytochemical, and cytogenetic characterization of a
human epithelioid sarcoma cell line (RM-HS1). J Natl Cancer
Inst 1987; 78: 7±18.
55. Quezado M. M. Alleli loss on chromosome 22q in epithelioid
sarcomas. Hum Pathol 1998; 29: 604±608.
56. Bolen J. W., Hammar S. P., McNutt M. A. Reactive and neoplastic
serosal tissue. A light-microscopic, ultrastructural, and
immunocytochemical study. Am J Surg Pathol 1986; 10: 43±47.
57. Sreekantaiah C., Ladanyi M., Rodriguez E. et al. Chromosomal
aberrations in soft tissue tumors. Relevance to diagnosis,
classification, and molecular mechanism. Am J Pathol 1994; 144:
1121±1134.
58. Sweet D. E., Vinh T. N., Devaney K. Cortical osteofibrous dysplasia
of long bone and its relationship to adamantinoma: A
clinicopathologic study of 30 cases. Am J Surg Pathol 1992; 16:
282±290.
59. Hazelbag H. M., Van den Broek L. J. C. M., Fleuren G. J. et al.
Distribution of extracellular matrix components in
adamantinoma of long bones suggests fibrous-to-epithelial
transformation. Hum Pathol 1997; 28: 183±188.
60. Dehner L. P. Primitive neuroectodermal tumor and Ewing's
sarcoma. Am J Surg Pathol 1993; 17: 1±13.
61. Gerharz C. D., Moll R., Ramp U. et al. Multidirectional
differentiation in a newly established human epithelioid sarcoma
cell line GRU1 with coexpression of vimentin cyto-keratin and
neurofilament proteins. Int J Cancer 1990; 45: 143±152.
62. Lawson C. W., Fisher C., Gatter K. C. An immunohistochemical
study of differentiation in malignant fibrous histiocytoma.
Histopathol 1987; 11: 375±383.
63. Noguera R, Navarro S, Cremades A. et al. Translocation 18 in a
biphasic synovial sarcoma with morphologic feature of neural
differentiation. Diagn Mol Pathol 1998; 7: 16±23.
64. Molenaar W. M., Muntinghe F. L. H. Expression of neural
cell adhesion molecules and neurofilament protein isoforms
in Ewing's sarcoma of bone and soft tissue sarcomas other
than rhabdomyosarcoma. Hum Pathol 1999; 30: 1207±1212.
664 Zhang et al.
Medical Hypotheses (2001) 57(5), 655±666
&
2001 Harcourt Publishers Ltd
65. Ordo
Ânez N. G. Desmoplastic small round cell tumor II. An
ultrastructural and immunohistochemical study with emphasis
on new immunohistochemical markers. Am J Surg Pathol 1998;
22: 1314±1327.
66. McManus A. P., Gusterson B. A., Pinkerton C. R. et al. The
molecular pathology of small round-cell tumors±relevance to
diagnosis, prognosis, and classification. J Pathol 1996;
178: 116±121.
67. Katz R. L., Quezado M., Senderowtcz A. M. et al. An
intra-abdominal small round cell neoplasm with features of
primitive neuroectodermal and desmoplastic round
cell tumor and a EWS/FLI1 fusion transcript. Hum Pathol 1997;
28: 502±509.
68. Fukunaga M., Nomura K., Endo Y. et al. Carcinosarcoma of the
uterus with extensive neuroectodermal differentiation.
Histopathol 1996; 29: 565±570.
69. Ladanyi M. The emerging molecular genetics of sarcoma
translocations. Diagn Mol Pathol 1995; 4: 162±173.
70. May W. A., Lessnick S. L., Braun B. S. et al. The Ewing's
sarcoma EWS/FLI-1 fusion gene encodes a more potent
transcriptional activator and is a more powerful
transforming gene than FLI-1. Mol Cell Biol 1993; 13:
7393±7398.
71. May W. A., Gishizky M. L., Lessnick S. L. et al. Ewing's sarcoma
11;22 translocation produces a chimeric transcription factor that
requires the DNA-binding domain encoded by FLI-1 for
transformation. Proc Natl Acad Sci USA 1993; 90: 5752±5756.
72. Teitell M. A., Thompson A. D., Sorensen P. H. B. etal. EWS/ET Sfusion
genes induce epithelial and neuroectodermal differentiation in
NIH 3T3 fibroblasts. Lab Invest 1999; 79:
1535±1543.
73. Sorensen P. H. B., Shimada H., Liu X. F. et al. Biphenotypic
sarcomas with myogenic and neural differentiation express the
Ewing's sarcoma EWS/FLI-1 fusion gene. Cancer Res 1995; 55:
1385±1392.
74. Thorner P., Squire J., Chilton-MacNeill S. et al. Is the EWS/FLI-I
fusion transcript specific for Ewing's sarcoma and peripheral
primitive neuroectodermal tumor? A report of four cases showing
this transcript in a wider range of tumor types. Am J Pathol 1996;
148: 1125±1138.
75. Dehner L. P. Peripheral and central primitive neuroectodermal
tumors. A nosologic concept seeking a consensus. Arch Pathol
Lab Med 1986; 110: 997±1005.
76. FruÈ hwald M. C., O'Dorisio M. S., Fleitz J. et al. Vasoactive intestinal
peptide (VIP) and VIP receptors: gene expression and growth
modulation in medulloblastoma and other central primitive
neuroectodermal tumors of childhood. Int J Cancer 1999; 81:
165±173.
77. Enjoji M., Hashimoto H., Tsuneyoshi M. et al. Malignant
fibrous histiocytoma. A clinicopathologic study of 130 cases.
Acta Pathol Jpn 1980; 30(5): 727±741.
78. Hoffman M. A., Dickersin G. R. Malignant fibrous histiocytoma.
An ultrastructural study of elevan cases. Hum Pathol 1983; 14:
913±922.
79. Lawson C. W., Fisher C., Gatter K. C. An immunohistochemical
study of differentiation in malignant fibrous histiocytoma.
Histopathol 1987; 11: 375±393.
80. Hirose T., Kudo E., Hasegawa T. et al. Expression of intermediate
filaments in malignant fibrous histiocytoma. Hum Pathol 1989;
20: 871±877.
81. Miettinen M., Soini Y. Malignant fibrous histiocytoma.
Hetergeneous patterns of intermediate filament proteins by
immunohistochemistry. Arch Pathol Lab Med 1989; 113:
1363±1366.
82. Rosenberg A. E., O'Counell J. X., Dickersin G. R. et al. Expression of
epithelial markers in malignant fibrous histiocytoma of the
musculoskeletal system. An immunohistochemical and electron
microscopic study. Hum Pathol 1993; 24: 284±293.
83. Bevis C. C. and Croker B. P. Bone and soft tissue neoplasms:
Contrasts and comparison by immunoperoxidase staining for
leukocyte antigens. Lab Invest 1985; 52: 7A±8A.
84. Roholl P. J. M., Kleyne J., van Unnik J. A. M. Characterization of
tumor cells in malignant fibrous histiocytomas and other soft
tissue tumors, in comparison with malignant histiocytes. II.
Immunoperoxidase study on cryostat sections. Am J Pathol 1985;
121: 269±274.
85. Roholl P. J. M., Kleijne J., van Basten C. D. H. et al. A study to
analyze the origin of tumor cells in malignant fibrous
histiocytomas. A multiparametric characterization. Cancer 1985;
56: 2809±2815.
86. Wood G. S., Beckstead J. H., Turner R. R. et al. Malignant fibrous
histiocytoma tumor cells resemble fibroblasts. Am J Surg Pathol
1986; 10: 323±335.
87. Brecher M. E., Franklin W. A. Absence of mononuclear phagocyte
antigens in malignant fibrous histiocytoma. Am J Clin Path 1986;
86: 344±348.
88. Iwasaki H., Isayama T., Ohjimi Y. et al. Malignant fibrous
histiocytoma. A tumor of facultative histiocytes showing
mesenchymal differentiation in cultured cell lines. Cancer 1992;
69: 437±447.
89. Shirasuna K., Sugiyama M., Miyazaki T. Establishment and
characterization of neoplastic cells from a malignant fibrous
histiocytoma. A possible stem cell line. Cancer 1985; 55:
2521±2532.
90. Strauchen J. A., Dimitriu-Bona A. D. Malignant fibrous
histiocytoma. Expression of monocyte/macrophage
differentiation antigens detected with monoclonal antibodies.
Am J Pathol 1986; 124: 303±309.
91. Binder S. W., Said J. W., Shintaku P. et al. A histiocyte-specific
marker in the diagnosis of malignant fibrous histiocytoma. Use
of monoclonal antibody KP-1 (CD68). Am J Clin Pathol 1992;
97: 759±763.
92. Yumoto T., Morimoto K. Experimental approach to fibrous
histiocytoma. Acta Pathol Jpn 1980; 30: 767±778.
93. Ghadially F. N., Roy S. Experimentally produced synovial
sarcomas. Cancer 1966; 19: 1901±1908.
94. Sakamoto K. Malignant fibrous histiocytoma induced by
intra-articular injection of 9,10-dimethyl-1, 2-benzanthracene
in the rat. Pathological and enzyme histochemical studies.
Cancer 1986; 57: 2313±2322.
95. Katenkamp D., Neupert G. Experimental tumors with features
of malignant fibrous histiocytomas. Light microscopic and
electron microscopic investigations on tumors produced by cell
transplantation of an established fibrosarcoma cell line.
Exp Pathol 1982; 22: 11±27.
96. Krawisz S. R., Florine D. L., Scott R. E. Differentiation of
fibroblast-like cells into macrophages. Cancer Res 1981;
41: 2891±2899.
97. Snover D. C., Sumner H. W., Dehner L. P. Variability of histologic
pattern in recurrent soft tissue sarcomas originally diagnosed
as liposarcoma. Cancer 1982; 49: 1005±1015.
98. Hashimoto H., Daimura Y., Tsuneyoshi M. et al. Soft tissue
sarcoma with additional anaplastic components. A
clinicopathologic and immunohistochemical study of 27 cases.
Cancer 1990; 66: 1578±1589.
99. Meis J. M. `Differentiation' in bone and soft-tissue tumors. A
histological indicator of tumor progression. Pathol Annual 1991;
26: 37±62.
Transdifferentiation of neoplastic cells 665
&
2001 Harcourt Publishers Ltd Medical Hypotheses (2001) 57(5), 655±666
100. Ro J. Y., Chen J. L., Lee J. S. et al. Sarcomatoid carcinoma of the
lung: Immunohistochemical and ultrastructural studies of 14
cases. Cancer 1992; 69: 376±386.
101. Nakajima M., Kasai T., Hashimoto H. et al. Sarcomatoid
carcinoma of the lung. A clinicopathologic study of 37 cases.
Cancer 1999; 86: 608±616.
102. Ro T. Y., Ayala A. G., Sella A. et al. Sarcomatoid renal cell
carcinoma: A clinicopathologic study of 42 cases. Cancer 1987;
59: 516±526.
103. Bertoni F., Ferri C., Benati A. et al. Sarcomatoid carcinomas of the
kidney. J Urol 1987; 137: 25±28.
104. Shannon R. L., Ro J. Y., Grignon D. J. et al. Sarcomatoid
carcinoma of the prostate. A clinicopathologic study of 12
patients. Cancer 1992; 69: 2676±2682.
105. Wargotz E. S., Norris H. J. Metaplastic carcinoma of the breast III
Carcinosarcoma. Cancer 1989; 64: 1490±1499.
106. Blom P. G., Stenwig A. E., Sobel H. J. et al. Intriguing cases. Case 6.
Ultrastruct Pathol 1983; 5: 307±313.
107. Ng H. K., Poon W. S. Gliosarcoma of the posterior fosa with
features of a malignant fibrous histiocytoma. Cancer 1990; 65:
1161±1166.
108. Chan J. K. C., Buchanan R., Fletcher C. D. M. Sarcomatoid variant
of anaplastic large-cell Ki-1 lymphoma. Am J Pathol 1990; 14:
983±986.
109. Orndal C., Rydholm A., Willen H. et al. Cytogenetic intratumor
heterogeneity in soft tissue tumors. Cancer Genet Cytogenet
1994; 78: 127±137.
110. Donner L. R. Cytogenetics of tumors of soft tissue and bone.
Cancer Genet Cytogenet 1994; 78: 115±126.
111. Mandahl N., Heim S., Willen H. et al. Characteristic karyotypic
anomalies identify subtypes of malignant fibrous histiocytoma.
Genes Chromosomes Cancer 1989; 1: 9±14.
112. Mandahl N., Heim S., Arheden K. et al. Rings, dicentrics, and
telomeric association in histiocytomas. Cancer Genet Cytogenet
1988; 30: 23±33.
113. Orndal C., Mandahl N., Rydholm A. et al. Supernumerary ring
chromosomes in five bone and soft tissue tumors of low or
borderline malignancy. Cancer Genet Cytogenet 1992; 60:
170±175.
114. Fletcher C. D. M., Akerman M., Cin P. D. et al. Correlation
between clinicopathological features and karyotype in
lipomatous tumors. A report of 178 cases from the
chromosomes and morphology (CHAMP) collaborative study
group. Am J Pathol 1996; 148: 623±629.
115. Rosai J., Akerman M., Cin P. D. et al. Combined morphologic and
karyotypic study of 59 atypical lipomaatous tumors. Evaluation
of their relationship and differential diagnosis with other
adipose tissue tumors. (A report of the CHAMP study group).
Am J Surg Pathol 1996; 20: 1182±1189.
116. Fletcher C. D. M., Cin P. D., de Wever I. et al. Correlation between
clinicopathological features and karyotype in spindle cell
sarcomas. A report of 130 cases from the CHAMP study group.
Am J Pathol 1999; 154: 1841±1847.
117. Ishikawa F: Breakthroughs and views. Telomer crisis, the driving
force in cancer cell evolution. Biochem Biophys Res Commun
1997; 230: 1±6.
118. Genberg M., Mark J., Hakalius L. et al. Origin and relationship
between different cell types in malignant fibrous histiocytoma.
Am J Pathol 1989; 135: 1185±1196.
119. Antonescu C. R., Kawai A., Leung D. H. et al. Strong association
of SYT-SSX fusion type and morphologic epithelial
differentiation in synovial sarcoma. Diagn Mol Pathol 2000;
9: 1±8.
120. Lengauer C., Kinzler K. W., Volgelstein B. Genetic instabilities in
humans. Nature 1998; 396: 643±649.
121. Pilotti S., Torre G. D., Lavarino C. et al. Distinct mdm2/p53
expression patterns in liposarcoma subgroups_implications for
different pathogenetic mechanisms. J Pathol 1997; 181: 14±24.
122. Hisaoka M., Morimitsu Y., Hashimoto H. et al. Retroperitoneal
liposarcoma with combined well-differentiated and myxoid
malignant fibrous histiocytoma-like myxoid areas. Am J Surg
Pathol 1999; 23: 1480±1492.
123. Damjanov I. Differentiation and transdifferentiation of normal
and neoplastic cells. Int J Dev Biol 1996; Suppl: 1: 63S.
666 Zhang et al.
Medical Hypotheses (2001) 57(5), 655±666
&
2001 Harcourt Publishers Ltd
... Loss of the differentiated phenotype is a main characteristic of the neoplastic cell. In many cell types, it occurs as a result of both loss of expression of differentiation markers and acquisition of phenotypic characteristics of a distinctly different cell type (transdifferentiation; Zhang & Xie 1994). In the thyroid follicular cell (TFC), the reduction/loss of expression of genes devoted to the synthesis of thyroid hormones occurring during transformation has been extensively described (Brabant et al. 1991, Ros et al. 1999, Schlumberger et al. 2007). ...
... In the thyroid follicular cell (TFC), the reduction/loss of expression of genes devoted to the synthesis of thyroid hormones occurring during transformation has been extensively described (Brabant et al. 1991, Ros et al. 1999, Schlumberger et al. 2007). Thus, sodium– iodide symporter (NIS), thyroperoxidase (TPO), thyroglobulin (Tg), and thyrotropin receptor (TSHR) gene expressions are progressively reduced or lost when the thyrocyte acquires more aggressive characteristics (Zhang & Xie 1994). Moreover, even expression of thyroid-specific transcription factors, such as TTF-1, HEX, and PAX8, is also down-regulated during transformation of the TFC (Fabbro et al. 1994). ...
Article
Full-text available
Periostin is a mesenchyme-specific gene product, which acts as an adhesion molecule during bone formation and supports osteoblastic cell line attachment and spreading. However, periostin expression is activated in a large variety of epithelial human tumors and correlates with their aggressiveness. Knowledge of expression of periostin in thyroid tumors is still scanty. The aim of the present work was to investigate periostin expression in differentiated neoplasms of the thyroid and to correlate it with several clinical and molecular features of these tumors. Periostin expression was evaluated by quantitative PCR and immunohistochemistry in normal thyroid tissues, papillary thyroid carcinomas (PTCs), follicular thyroid carcinomas (FTCs), and follicular adenomas (FAs). Periostin mRNA levels were also evaluated in several thyroid tumor cell lines. PTCs show mean periostin mRNA levels significantly higher than corresponding normal tissues. In five PTCs, periostin mRNA values were at least 30-fold higher than corresponding normal tissues. Conversely, mean periostin mRNA levels of FTCs and FAs were similar to those of normal tissues. Consistent with mRNA studies, periostin was detectable by immunohistochemistry in cancerous epithelial cells only in several cases of PTCs but not in normal tissue, FTCs, and FAs. In PTCs, periostin mRNA levels positively correlate with extrathyroidal invasion, distant metastasis, and higher grade staging. A negative correlation between periostin and expression of some markers of the thyroid-differentiated phenotype (thyroglobulin, thyrotropin receptor) was also present in the PTCs. These results indicate that an increase in periostin gene expression is present in several PTCs, in which it appears as a marker of aggressiveness. Experiments in thyroid tumor cell lines indicate that high levels of periostin mRNA are due, at least in part, to the increase in periostin promoter activity.
... There are several lines of evidence that cancer classifications may not reflect tumour histogenesis but tumour differentiation (Gusterson et al., 2005; Gusterson, 2009; Gould, 1986; Zhang et al., 2001; Moinfar, 2008 ). Nevertheless, the proponents of the molecular classification for breast cancer envisaged that the molecular subgroups identified by microarray analysis would indeed have histogenetic implications (i.e. that the phenotype of a breast cancer would mirror that of its cell of origin) (Sorlie et al., 2001). ...
Article
Breast cancer is a heterogeneous disease, comprising multiple entities associated with distinctive histological and biological features, clinical presentations and behaviours and responses to therapy. Microarray-based technologies have unravelled the molecular underpinning of several characteristics of breast cancer, including metastatic propensity and histological grade, and have led to the identification of prognostic and predictive gene expression signatures. Furthermore, a molecular taxonomy of breast cancer based on transcriptomic analysis has been proposed. However, microarray studies have primarily focused on invasive ductal carcinomas of no special type. Owing to the relative rarity of special types of breast cancer, information about the biology and clinical behaviour of breast cancers conveyed by histological type has not been taken into account. Histological special types of breast cancer account for up to 25% of all invasive breast cancers. Recent studies have provided direct evidence of the existence of genotypic-phenotypic correlations. For instance, secretory carcinomas of the breast consistently harbour the t(12;15) translocation that leads to the formation of the ETV6-NTRK3 fusion gene, adenoid cystic carcinomas consistently display the t(6;9) MYB-NFIB translocation and lobular carcinomas consistently show inactivation of the CDH1 gene through multiple molecular mechanisms. Furthermore, histopathological and molecular analysis of tumours from conditional mouse models has provided direct evidence for the causative role of specific genes in the genesis of specific histological special types of breast cancer. Here we review the associations between the molecular taxonomy of breast cancer and histological special types, discuss the possible origins of the heterogeneity of breast cancer and propose an approach for the identification of novel therapeutic targets based on the study of histological special types of breast cancer.
... It could be hypothesized that COMP has a potential role in epithelial-mesenchymal transdifferentiation of pancreatic acini cells. However, this term usually describes the transition of tumor cells from an epithelia to a mesenchymal phenotype (36), as it has been observed for example in pancreatic cancer cells (37). It will be interesting to further analyze whether pancreatic acini undergo epithelial-mesenchymal transdifferentiation in the course of CP, whether COMP is directly or indirectly involved in this process, and whether this transdifferentiation plays a role in the development of the desmoplastic reaction in CP. ...
Article
Full-text available
Chronic pancreatitis (CP) is an inflammatory disease of the pancreas characterized histomorphologically by progressive development of fibrosis and atrophy of the pancreatic parenchyma. Cartilage oligomeric matrix protein (COMP) is a member of the thrombospondin (TSP) family of extracellular glycoproteins that is expressed in CP tissues. In the present study, we characterized COMP mRNA and protein expression in the normal pancreas, chronic pancreatitis, and pancreatic cancer tissues. 15 normal pancreatic tissues, 14 CP tissues and 14 pancreatic cancer tissues were analyzed by Northern blotting, Western blotting, in situ hybridization and immunohistochemistry. COMP mRNA and protein were detected at moderate to high levels in chronic pancreatitis tissues, at moderate levels in pancreatic cancer tissues, but at low levels in normal pancreatic tissues and in four pancreatic cancer cell lines. COMP mRNA signals and immunoreactivity were strongly present in the cytoplasm of degenerating acinar cells in CP tissues as well as in CP-like lesions in pancreatic cancer tissues. COMP protein was also present in the fibrotic tissue in CP. In contrast, COMP expression was weak to absent in the cytoplasm of cancer cells in pancreatic cancer tissues, and in ductal cells and islet cells in normal pancreatic tissues. COMP is preferentially expressed in degenerating acinar cells in CP and in CP-like areas in pancreatic cancer, suggesting a potential role of this gene in the course of acinar cell degeneration and dedifferentiation. COMP might thus serve as a marker for tissue destruction and disease activity in CP.
... These results may explain the divergent effects of Id2 in epithelial versus uveal melanoma cells and predict the existence of a strong activator of the E-box2 in uveal melanoma (Fig. 3D). This conclusion is consistent with previous work showing that suppression of E-cadherin expression in nonepithelial cells is maintained by constitutive repression of the promoter (24) and that Id2 suppresses epithelial differentiation in the neural crest lineage (25). Id2 itself does not contain a DNA-binding domain or interact directly with the promoter but rather regulates gene expression through inhibitory interactions with basic helix-loop-helix transcription factors (26). ...
Article
Full-text available
Microarray gene expression profiling is a powerful tool for generating molecular cancer classifications. However, elucidating biological insights from these large data sets has been challenging. Previously, we identified a gene expression-based classification of primary uveal melanomas that accurately predicts metastatic death. Class 1 tumors have a low risk and class 2 tumors a high risk for metastatic death. Here, we used genes that discriminate these tumor classes to identify biological correlates of the aggressive class 2 signature. A search for Gene Ontology categories enriched in our class-discriminating gene list revealed a global down-regulation of neural crest and melanocyte-specific genes and an up-regulation of epithelial genes in class 2 tumors. Correspondingly, class 2 tumors exhibited epithelial features, such as polygonal cell morphology, up-regulation of the epithelial adhesion molecule E-cadherin, colocalization of E-cadherin and beta-catenin to the plasma membrane, and formation of cell-cell adhesions and acinar structures. One of our top class-discriminating genes was the helix-loop-helix inhibitor ID2, which was strongly down-regulated in class 2 tumors. The class 2 phenotype could be recapitulated by eliminating Id2 in cultured class 1 human uveal melanoma cells and in a mouse ocular melanoma model. Id2 seemed to suppress the epithelial-like class 2 phenotype by inhibiting an activator of the E-cadherin promoter. Consequently, Id2 loss triggered up-regulation of E-cadherin, which in turn promoted anchorage-independent cell growth, a likely antecedent to metastasis. These findings reveal new roles for Id2 and E-cadherin in uveal melanoma progression, and they identify potential targets for therapeutic intervention.
Article
Drugs which elicit cell differentiation might have an important role in the treatment of leukemias and other neoplasias. Various chemotherapeutic agents promote leukemic cell differentiation. The HL-60 cell line is a useful model to study in vitro myeloid differentiation. Sublethal concentrations of 3-nitrobenzothiazolo[3,2-a]quinolinium (NBQ), an antitopoisomerase II drug, were given to HL-60 cells from one to five days to evaluate its capacity to induce differentiation. NBQ-induced HL-60 cells reduced nitroblue tetrazolium (NBT), increased MY-4 receptors, increased phagocytic activity and displayed the granulocytic morphology. Flow cytometric DNA analysis of NBQ-induced cells revealed an arrest in the G1 phase a reduction in the relative percentage of cells in S and G2+M phases. Our results suggest that NBQ induces an S-phase specific differentiation of HL-60 cells comparable to that previously described with dimethyl sulfoxide and retinoic acid. NBQ and its analogs, as differentiation inducers, may have potential utility as a novel therapeutic modality for leukemias.
Article
Pleomorphic adenomas of the parotid gland are benign tumours composed of epithelial and mesenchymal cells. The INK4a-ARF (CDKN2A) locus on chromosome 9p21 encodes two tumour suppressor proteins, p16(INK4a) and p14(ARF), which act as upstream regulators of the Rb-CDK4 and p53 pathways. To study the contribution of each pathway in pleomorphic adenomas, this study analysed alterations of p14(ARF), p16(INK4a), p53, and pRb in these tumours. After microdissecting the different histological components, 42 pleomorphic adenomas of the parotid gland were analysed for INK4a-ARF inactivation by DNA sequence analysis, methylation-specific PCR (MSP), restriction enzyme-related polymerase chain reaction (RE-PCR), mRNA expression, microsatellite analysis, and immunohistochemistry. In addition, microdeletion of p14(ARF) and p16(INK4a) were assessed by differential PCR. The status of p53 and Rb was examined by direct sequencing and immunohistochemistry. Using microdissection, it was possible to examine the tumour components, i.e. epithelial, mesenchymal, and transitional, separately after immunohistochemical identification. Methylation of p14(ARF) was found in 1/42 cases and alterations of p16(INK4a) occurred in 12/42 of pleomorphic adenomas, which correlated with loss of mRNA transcription. Microdeletions or specific mutations of either exon were not detected. Methylation was detected exclusively in the epithelial and transitional components and not within the mesenchymal part of the tumour. p53 mutations were detected in 4/42 adenomas, also occurring solely in the epithelial components of the tumours. pRb was detected immunohistochemically in 40/42 adenomas. In normal, corresponding parotid tissue, p14(ARF), p16(INK4a), p53, and pRb alterations were not observed. The observation that alterations of p14(ARF) and p16(INK4a), and also p53 mutations, occurred exclusively in the epithelial and transitional components of pleomorphic adenoma supports the theory that these areas are prone to malignant transformation to carcinoma in adenoma.
Article
Morules are a diagnostic clue to the cribriform-morular variant (C-MV) of papillary thyroid carcinoma, and are superficially similar to squamous metaplasia. In order to clarify the histogenesis of morules and differentiate them from squamous metaplasia, we immunohistochemically compared the morules in five cases of C-MV with squamous metaplasia in six cases of diffuse sclerosing variant (DSV) of papillary thyroid carcinoma. The squamous metaplastic cells were immunopositive for low- and high-molecular-weight cytokeratin, whereas the morular cells were negative or focally positive. Vimentin-positive cells were observed focally in the morules and squamous metaplasia, except for one case of CMV that showed intense positivity. The morular cells showed weak cytoplasmic positivity for beta-catenin, and the cell membrane was not highlighted. Some nuclei of the morular cells were also positive for this antibody. Beta-catenin was intensively positive along the cell membrane of the metaplastic cells, and did not react against the nuclei or cytoplasm. Bcl-2 was positive in the morular cells, but negative in the metaplastic cells. S-100 protein-positive dendritic cells were observed in the metaplastic nests, but not in the morules. We argue that morules appear in connection with nuclear and cytoplasmic aberrant localization of beta-catenin, and are not an early form of squamous metaplasia.
Article
The distinction between sarcomatoid carcinoma (SC) and bona fide sarcoma can be difficult using conventional immunohistochemical markers. Epithelial-mesenchymal transition (EMT) has been proposed as a histogenetic mechanism for the development of SC. Expression of selected markers of EMT (Twist and Slug) was compared with other markers of epithelial differentiation in SC and spindle cell sarcoma to determine the utility of these antigens in this differential diagnosis. Twenty-seven cases of SC (excluding those of gynecologic origin) were stained by immunohistochemistry for cytokeratins (AE1/AE3, 5D3, CK5/6, and 34betaE12), p63, claudin-1, claudin-7, epithelial cadherin, placental cadherin, epithelial cell adhesion molecule/epithelial-specific antigen, 14-3-3sigma, Twist, and Slug. A comparison group of 21 spindle or pleomorphic spindle cell sarcomas was also studied. Immunohistochemical stains were scored in a semiquantitative manner and subsequent exploratory analyses were performed using logistic regression and chi2 tests. Only cytokeratin AE1/AE3 specifically labeled SC in a statistically significant manner. Other epithelial-specific markers tested did not distinguish SC from sarcoma primarily owing to low sensitivity. However, when positive, immunostains such as CK5/6, membranous epithelial cadherin, and nuclear p63 may aid in the distinction of SC from sarcoma. EMT markers were expressed in most cases of both SC and sarcoma, and were not useful in making a differential diagnosis between these neoplasms.
Article
Epithelial-mesenchymal transition (EMT) is a phenotypic conversion that facilitates organ morphogenesis and tissue remodeling in physiological processes, such as embryonic development and wound healing. A similar phenotypic conversion is also detected in fibrotic diseases and neoplasia, and is associated with disease progression. EMT in cancer epithelial cells often seems to be an incomplete and bidirectional process. In this Review, we discuss the phenomenon of EMT as it pertains to tumor development, focusing on exceptions to the commonly held rule that EMT promotes invasion and metastasis. We also highlight the role of RAS-controlled signaling mediators, ERK1, ERK2 and phosphatidylinositol 3-kinase, as microenvironmental responsive regulators of EMT.
Article
Full-text available
Vasoactive intestinal peptide (VIP) is a neuromodulator and growth regulator in the developing nervous system. We analyzed 10 primitive neuroectodermal tumor (PNET) cell lines, 29 central PNET (cPNET) and 17 tumors of the Ewing's sarcoma/peripheral PNET family (ESFT) using reverse transcriptase‐polymerase chain reaction (RT‐PCR) and Southern hybridization. Each of the 10 cell lines and 86.2% of cPNET expressed mRNA for VIP receptor 1 (VIPR1) compared to 52.9% of ESFT. VIPR2 was expressed in 75.8% of cPNET, in 28.6% of ESFT and in all 10 cell lines. cPNET demonstrated high‐affinity binding of ¹²⁵I‐VIP on quantitative autoradiography and in competitive binding assays. VIP inhibited tumor cell proliferation in a dose‐dependent manner in 5 of 7 PNET cell lines. We conclude that VIPR1 and VIPR2 are highly expressed in cPNET and demonstrate that VIP is a growth modulator in these tumors. Int. J. Cancer 81:165–173, 1999. © 1999 Wiley‐Liss, Inc.
Article
The origin of malignant mixed Müllerian tumours (MMMTs) has long been debated, due to the indefinite relationship between epithelial and mesenchymal malignant cells. In order to obtain insight into the clonal relationship between the two components of these tumours, molecular genetic changes were investigated at the level of loss of heterozygosity (LOH) in both cells types. LOH was studied in a series of six cases with 74 polymorphic microsatellite markers mapping to 19 different chromosomes. The epithelial and the mesenchymal neoplastic cells were separately microdissected from formalin‐fixed, paraffin‐embedded tissue, prior to DNA isolation. LOH was observed for 35 different markers mapping to chromosomes 3, 6, 8, 11, 15, 16, 17, 18, 21, and X. The most frequently involved chromosomes were 17p, 17q, 11q, 15q, and 21q. LOH was observed in five out of six cases and identical alleles were lost in the epithelial and in the mesenchymal cells. No genetic differences were observed between the two cell types for any of the informative markers. Immunohistochemistry (IHC) and TP53 mutation analysis revealed involvement of TP53 in all cases. Mutations were identified in five MMMTs. In four tumours, of which three had a missense mutation, strong nuclear staining for p53 was observed. In the remaining two cases, the mutation resulted in a stop codon, with no nuclear staining for p53 by IHC. The results support a monoclonal origin of MMMTs, with the absence of genetic changes uniquely associated with either of the phenotypes. The latter finding is compatible with current opinion that these neoplasms should be considered as metaplastic carcinomas and supports the conversion hypothesis. © 1997 John Wiley & Sons, Ltd.
Article
Scatter Factor, also known as Hepatocyte Growth Factor (SF/HGF), has pleiotropic functions including direct control of cell-cell and cell-substrate adhesion in epithelia. The subcellular localization of the SF/HGF receptor is controversial. In this work, the cell surface distribution of the SF/HGF receptor was studied in vivo in epithelial tissues and in vitro in polarized MDCK monolayers. A panel of monoclonal antibodies against the beta chain of the SF/HGF receptor stained the basolateral but not the apical surface of epithelia lining the lumen of human organs. Radiolabeled or fluorescent-tagged anti-receptor antibodies selectively bound the basolateral cell surface of MDCK cells, which form a polarized monolayer sealed by intercellular junctions, when grown on polycarbonate filters in a two-chamber culture system. The receptor was concentrated around the cell-cell contact zone, showing a distribution pattern overlapping with that of the cell adhesion molecule E-cadherin. The basolateral localization of the SF/HGF receptor was confirmed by immunoprecipitation after domain selective cell surface biotinylation. When cells were fully polarized the SF/HGF receptor became resistant to non-ionic detergents, indicating interaction with insoluble component(s). In pulse-chase labeling and surface biotinylation experiments, the newly synthesized receptor was found exclusively at the basolateral surface. We conclude that the SF/HGF receptor is selectively exposed at the basolateral plasma membrane domain of polarized epithelial cells and is targeted after synthesis to that surface by direct delivery from the trans-Golgi network.
Article
Many of the major solid, malignant tumors of childhood have histologic similarities that reflect their dysembryonic and primitive features. One subset of these neoplasms, Ewing's sarcoma (ES) and primitive neuroectodermal tumor (PNET), presents primarily in the bone and soft tissues. Both tumor types were reported at a time and date well before the advent of electron microscopy and immunohistochemistry. Opposition to ES and PNET as distinctive entities developed and persisted because these tumors were considered incompletely documented examples of metastatic neuroblastoma or malignant lymphoma. General acceptance of ES as a unique tumor type occurred well before the PNET had been fully defined and characterized. Once these neoplasms had joined the other round cell neoplasms, the quest for the histogenesis was pursued, but the results were frustratingly inconclusive, especially for ES. Because of the resemblance of the PNET to classic neuroblastoma, the neural crest was regarded as the most likely progenitor. With the recognition of osseous PNET, extraosseous ES, and a shared cytogenetic abnormality between ES and PNET, more recent speculation has focused on the possibility that these presumably separate neoplasms are closely related histogenetically without directly answering the question of histogenesis. Despite the likely common progenitorship of ES and PNET, the latter neoplasm is seemingly the more aggressive. Although melanotic neuroectodermal tumor of infancy, intra-abdominal desmoplastic small cell tumor, and polyphenotypic small cell tumors have some overlapping microscopic and immunohistochemical features with PNET, their relationship to ES-PNET has otherwise not been resolved.
Article
Malignant mixed tumors (carcinosarcomas) are examples of unusual neoplasms whose occurrences have been observed in increasingly diverse sites but whose pathogenesis remains a complete mystery. Two antithetical hypotheses that have been advanced to explain the histogenesis of these tumors include the convergence hypothesis, which proposes an origin from two or more stem cells (multiclonal hypothesis), and the divergence hypothesis, which proposes an origin from a single totipotential stem cell that differentiates into separate epithelial and mesenchymal directions (monoclonal hypothesis). To test these hypotheses, a novel strategy for the determination of clonality from as few as 100 tumor cells obtained by enzymatic digestion of either fresh or formalin-fixed, paraffin-embedded tissues and cell sorting was used that exhibited the polymerase chain reaction (PCR) in amplifying a 511-bp region located within the first intron of the human hypoxanthine phosphoribosyl transferase gene, a site that contains inactive X chromosomal obligately methylated HpaII/MspI sites and single-base allelic polymorphisms in 5% females. Carcinoma cells gated on the basis of fluorescein isothiocyanate (FITC)-anti-cytokeratin and sarcoma cells gated on the basis of FITC-anti-vimentin or FITC-anti-desmin were sorted to homogeneity on FACSTAR and then subjected to genomic DNA extraction and Hpa II digestion before PCR amplification and subsequent analysis of the product on denaturing gradient gel electrophoresis. The comigrations of the single homoduplexes generated from both the carcinoma cells and sarcoma cells in six different malignant mixed tumors obtained from four different organs indicated clonal identity and monoclonality in all cases. These findings of monoclonality were confirmed independently by two other methods of clonality determination. The findings of a monoclonal origin of carcinosarcomas support the single totipotential stem-cell-divergence hypothesis.
Article
Malignant change in a benign eccrine spiradenoma is rare. Only 13 such cases have been previously reported. Two further cases are described herein, both of which were carcinosarcomas and arose in middle-aged women. In each case continuity was demonstrated between the benign and malignant epithelial components and the sarcomatous element. In case 1 the sarcomatous component showed osteocartilaginous and rhabdomyoblastic differentiation, whereas in case 2 only osteosarcoma was evident. To our knowledge, carcinosarcoma has not been previously described arising in eccrine spiradenoma. The literature regarding malignant eccrine spiradenoma is reviewed.