Vous êtes sur la page 1sur 22

Cancer and Metastasis Reviews 22: 337358, 2003.

# 2003 Kluwer Academic Publishers. Manufactured in The Netherlands.

Src family kinases in tumor progression and metastasis


Justin M. Summy and Gary E. Gallick*
The University of Texas M.D. Anderson Cancer Center, Department of Cancer Biology, Houston TX 77030,
USA

Key words: Src, protein tyrosine kinase, tumor progression, angiogenesis, cell survival
Summary
The Src family of non-receptor protein tyrosine kinases plays critical roles in a variety of cellular signal
transduction pathways, regulating such diverse processes as cell division, motility, adhesion, angiogenesis,
and survival. Constitutively activated variants of Src family kinases, including the viral oncoproteins v-Src
and v-Yes, are capable of inducing malignant transformation of a variety of cell types. Src family kinases,
most notably although not exclusively c-Src, are frequently overexpressed and/or aberrantly activated in a
variety of epithelial and non-epithelial cancers. Activation is very common in colorectal and breast cancers,
and somewhat less frequent in melanomas, ovarian cancer, gastric cancer, head and neck cancers,
pancreatic cancer, lung cancer, brain cancers, and blood cancers. Further, the extent of increased Src family
activity often correlates with malignant potential and patient survival. Activation of Src family kinases in
human cancers may occur through a variety of mechanisms and is frequently a critical event in tumor
progression. Exactly how Src family kinases contribute to individual tumors remains to be dened
completely, however they appear to be important for multiple aspects of tumor progression, including
proliferation, disruption of cell/cell contacts, migration, invasiveness, resistance to apoptosis, and
angiogenesis. This review details the evidence for Src family activation in human tumors, and
emphasizes possible consequences to tumor progression. Given the ability of Src and its family members
to participate in so many aspects of tumor progression and metastasis, Src family kinases are attractive
targets for future anti-cancer therapeutics.

Introduction
In 1911, the young pathologist, Peyton Rous,
isolated from a chicken the virus that has
continued to bear his name, Rous Sarcoma Virus
[1]. While Rous was able to fulll all of Kochs
postulates with respect to demonstrating that the
virus was an etiologic agent for sarcomas in these
birds, the full signicance of the work went
unrecognized for decades, and it was not until
1966 that he was to win the Nobel Prize in
Medicine for his accomplishments. As is now
obvious, Rous Sarcoma Virus was the archetypal
retrovirus, which collectively harbor transforming
* Corresponding author.
E-mail: ggallick@mdanderson.org

genes that were to revolutionize our understanding


of signal transduction and cancer biology. Rous
Sarcoma Virus and its transforming gene termed
Src (originally because it induced Sarcomas in
infected birds) have been at the vanguard of
advances in our understanding of signal transduction and malignant transformation. By an elegant
strategy using tumor bearing rabbit serum [2], Src
was the rst transforming protein isolated.
Through then very tedious molecular approaches,
Bishop and Varmus [3] demonstrated that the viral
Src gene (v-src) had as its progenitor a normal
cellular gene (proto-oncogene) c-src, an experiment largely responsible for the awarding of the

338

Summy and Gallick

Nobel Prize in Medicine to them in 1989. Src was


also discovered to be the rst gene product with
intrinsic protein tyrosine kinase activity [4].
With such an illustrious history, one would
expect alterations in Src would also play a leading
role in understanding how aberrant activation of
proto-oncogenes was important in human tumorigenesis and tumor progression. However, such
has not been the case. Because Src is rarely
mutated in human tumors, and amplications
and rearrangements of the gene in tumors are even
rarer, establishing a clear role for Src in human
cancers has proven difcult. Nevertheless, as our
knowledge of signal transduction has increased,
and the central role of Src in many signaling
processes continues to be discerned, the concept
has become better established that abnormal
regulation of Src and its family members are
critical in many types of tumors. This review will
rst briey describe Src structure and function,
then focus on the best studied tumors in which
aberrant expression or activity of Src may be
critical to the disease.

The Src family of protein tyrosine kinases


The Src family kinases (SFKs) comprise a subclass of membrane-associated non-receptor tyrosine kinases that are involved in a variety of
cellular signal transduction pathways. There are
nine members of the Src family, including c-Src,
c-Yes, Fyn, Lyn, Lck, Hck, Blk, Fgr, and Yrk.
Most Src family kinases are expressed primarily in
cells of hematopoietic origin. C-Src, c-Yes, and
Fyn, however, display a much more ubiquitous
pattern of expression, with particularly high levels
in platelets, neurons, and some epithelial tissues
[5]. Src family kinases are activated in response to
cellular signals that promote proliferation, survival, motility, and invasiveness, including activation of cytokine receptors, receptor protein
tyrosine kinases, g-protein coupled receptors, and
integrins [5]. Their involvement in these normal
cellular signaling pathways, coupled with the
ability of constitutively active Src family kinases
to induce cell transformation, has sparked a keen
interest in the role Src family kinases may play in
the onset or progression of human cancers.
Indeed, c-Src and c-Yes are overexpressed and/or

hyper-activated in a variety of human epithelial


cancers, while the primarily hematopoietic-specic
SFKs may be involved in the growth and
progression of some leukemias and lymphomas.

Src family kinase structure and regulation


The ability of the avian viral oncoproteins v-Src
and v-Yes to induce broblast cell transformation
suggests that their cellular counterparts, c-Src and
c-Yes, have the potential to contribute to human
carcinogenesis. V-Src and v-Yes are encoded by
avian retroviruses and are capable of inducing
sarcomas in chickens and transforming chicken
embryo broblast cells in culture [6,7]. To understand how these proteins are able to induce cell
transformation, it is important to understand the
functional domain architecture shared by all Src
family kinases, and the role of these domains in
both regulating tyrosine kinase activity and
recruiting additional proteins into signaling complexes. These aspects have been reviewed extensively elsewhere [5,8], and will only be discussed
briey here.
All Src family kinases are comprised of an
amino terminal membrane localization signal, also
known as the Src homology 4 or SH4 domain, a
poorly conserved unique domain, an SH3 domain,
SH2 domain, tyrosine kinase domain, and a
regulatory sequence [8]. Src family kinases are
kept inactive through an elegant series of intramolecular interactions that restrict the accessibility
of the kinase domain active site for ATP and
substrates. The inactive or closed conformation is
attained via phosphorylation of a conserved
tyrosine residue (Y527 in avian c-Src), near the
carboxy terminal tail of the protein, which forms
an intramolecular interaction with the SH2
domain [9,10]. Working cooperatively with the
SH2/tail interaction is an association between the
SH3 domain and the short stretch of amino acids
linking the SH2 and kinase domains [9,10].
Together, these intramolecular interactions not
only limit accessibility of the kinase domain but
also the binding surfaces of the SH3 and SH2
domains, thus further limiting the potential for
these proteins to participate in cellular signaling.
The viral proteins v-Src and v-Yes, through loss
of sequence in their carboxy-terminal tails, are no

Src family kinases in tumor progression and metastasis

longer able to be regulated through intramolecular


interactions and thus are constitutively active and
transformation-competent [6,7]. Mutation of Y527
in c-Src to phenylalanine also results in an active
and transformation competent protein [1113].
However, although there has been at least one
report of a c-Src activating mutation present in
human colon cancer [14], genomic mutation is not
the predominant mechanism of SFK activation in
human cancers. In the absence of mutation, kinase
activity may be augmented through displacement
of the intramolecular interactions via higher
afnity binding between the SH3 and SH2
domains and their cellular ligands and/or through
dephosphorylation of the carboxy-terminal regulatory tyrosine, normally phosphorylated by the
C-terminal Src kinase (Csk) and closely related
kinases [8,15,16]. Src family kinase activation
arises through multiple mechanisms, differing not
only between different cancers, but also within a
particular subclass of cancers. Regardless of the
mechanism, the constitutively high specic activity
of SFKs in a large number of human tumors, even
in the absence of activating mutations suggest that
they may be important in the malignant process.
Further, as discussed throughout this review, the
rarity of activating mutations suggest that SFK
activation may be more critical for tumor progression than tumor initiation.
How then do SFKs contribute to the progression of cancers? The available evidence would
suggest that they are capable of promoting several
aspects of tumor progression and metastasis. The
hallmarks of a metastatic cancer are multi-fold
[17]. Regulation of Src family members has been
implicated in several of these, including aberrant
proliferation; increased motility and invasiveness;
resistance to apoptosis; and angiogenesis. Not
surprisingly, these are normally regulated, in part
by Src family members during development,
differentiation, and homeostasis. In this review,
we will focus on how their aberrant regulation
leads to tumor progression in a variety of human
tumors.

Colorectal carcinoma
The activation and functions of Src family kinases
have been more extensively studied in colon cancer

339

than in any other human cancer. The initial studies


of c-Src in colon cancer were largely correlative
and indicated that c-Src protein levels and/or
kinase activity are frequently elevated in colon
carcinomas relative to normal colonic mucosa. In
1986, Rosen and co-workers [18] discovered
elevated c-Src kinase activity in all colon cancer
cell lines studied. Similarly, Bolen et al. [19]
reported elevated c-Src specic kinase activity in
colon carcinoma tissues and cell lines, relative to
normal colonic mucosa. The elevated kinase
activity was not due solely to elevated c-Src
protein expression levels. Furthermore, the elevated kinase activity appeared to be the result of a
post-translational modication, as in vitro translated mRNA from colon tumor cells did not result
in c-Src protein with elevated kinase activity
relative to that from normal colonic mucosa [19].
Similar results were obtained by Cartwright et al.
[20].
Later studies demonstrated that c-Src kinase
activity positively correlated with the malignant
potential of the cells, providing evidence that c-Src
activation contributes to the progression of colon
cancer toward the metastatic phenotype. In colon
cancer, evidence suggests that Src activation might
contribute both to early and later stages of tumor
progression. In 1990, Cartwright and colleagues
determined that c-Src kinase activity was elevated
in adenomatous polyps relative to normal mucosa,
and that adenomas with the greatest malignant
potential, having a villous structure and severe
dysplasia, displayed the highest kinase activity
[21]. Further studies from this laboratory determined that c-Src kinase activity was elevated in
premalignant ulcerative colitis lesions, and that,
again, those lesions with the greatest dysplasia and
thus the greatest potential for progression to
carcinoma, displayed the highest levels of c-Src
activity [22].
Results obtained by Weber et al. [23] indicated
that c-Src activity is highest in moderate to well
differentiated colonic lesions, while poorly differentiated carcinomas displayed c-Src kinase activity
similar to normal colonic mucosa. These studies
have not been pursued further to determine
whether cell composition (tumor versus stromal
cells; necrosis, etc.) might have contributed to this
observation. Talamonti and co-workers [24], in
1993, compared levels of c-Src activity from

340

Summy and Gallick

colonic polyps, primary tumor lesions, and liver


metastases. In these studies it was observed that
Src kinase activity was elevated in pre-malignant
polyps, higher still in primary colonic lesions, and
highest in liver metastases. These results again
provide correlative evidence that c-Src activity
may promote progression toward the metastatic
phenotype. Studies by Termuhlen et al. [25]
conrmed these results and indicated that the
increased c-Src activity in hepatic metastases was
specic for colon carcinoma, as non-colon-derived
hepatic metastases displayed little elevation in cSrc activity. However, increased c-Src activity was
not limited to hepatic metastases, as extrahepatic
colonic metastases displayed elevated c-Src activity, which was on average, 5.7 fold higher than
that detected in hepatic metastases [25]. These
results indicate that c-Src activity may contribute
to the metastatic progression of colonic metastases, regardless of the site of metastasis. Interestingly, although the majority of the available data
indicate that c-Src kinase activity increases with
increasing metastatic potential in colon carcinomas, there are some reports that indicate that c-Src
activation occurs primarily in the early stages of
colon cancer progression. Sakai et al. [26] reported
that adenomas displayed stronger immunohistochemical staining for active c-Src than did more
advanced adenocarcinomas. Iravani et al. [27], in
1998, reported that c-Src expression and kinase
activity were highest in adenomatous colon tissue,
relative to normal colonic mucosa, primary
carcinoma lesions, or synchronous metastatic
lesions. These results suggest that c-Src activation
contributes to cancerous progression primarily in
the early stages of carcinoma formation. The
reasons for these discrepancies are unclear; however, it may simply indicate that c-Src activation,
while occurring frequently in the progression of
colonic epithelium toward the metastatic phenotype, is not necessarily a requirement at every stage
of progression. Nevertheless, the majority of
studies using human tissue suggest that c-Src
activation likely contributes to malignant progression of colon cancer. Perhaps the strongest
correlative evidence that Src contributes to tumor
progression was obtained when Allgayer et al. [28]
recently reported that elevated c-Src activity served
as an independent indicator of prognosis at all
stages of carcinoma. The ability of Src activation to

predict tumor cell survival has future implications


in therapeutic treatment of colon cancer.
Although c-Src is frequently activated in human
colon cancer, it is not the only Src family kinase
whose activation corresponds with malignant
progression. C-Yes is also frequently activated in
colon carcinoma. Park and colleagues, in 1993,
demonstrated that c-Yes kinase activity was
elevated in three of ve colon cancer cell lines
and 10 of 21 primary colon carcinomas, relative to
normal colonic mucosa [29]. It was reported in this
study that the elevated c-Yes kinase activity was
largely attributable to elevated protein levels. Pena
et al. [30] reported in 1995 that, similar to c-Src, cYes kinase activity was also elevated in premalignant lesions of the colon and that the degree of cYes activation correlated positively with the likelihood of malignant progression. Adenomas at
highest risk for progression to cancer, those greater
than 2 cm in diameter with severe dysplasia and a
villous architecture, displayed the highest level of
c-Yes kinase activity [30]. In 1996, Alexander and
colleagues demonstrated that c-Yes message levels
were elevated in inammatory bowel disease (IBD)
lesions, relative to normal colon [31]. C-Yes
message levels were higher in IBD patients who
were resected for colon cancer than those who were
not [31]. Interestingly, Han et al. [32] reported that,
similar to c-Src, c-Yes is frequently activated in
colon cancer liver metastases. However, rarely are
both kinases found to be activated in the same
tumor and c-Yes activation generally corresponds
with a worse clinical prognosis than does c-Src
activation in liver metastases.
Given the high homology between c-Src and cYes, it is frequently assumed that they perform
redundant functions. While there is considerable
evidence to support this notion, several studies in
recent years have also indicated specicity in
signaling between c-Src and c-Yes. Thus the
question arises as to whether c-Src and c-Yes are
regulated in the same fashion in colon carcinoma
and whether or not they perform unique or
redundant functions. Unfortunately, very little is
known about the role of c-Yes in colon. It is
known, however, that cells from c-src  /  mice
are decient in several processes that are dependent on the dynamic regulation of the actin
cytoskeleton, including motility, cell spreading,
membrane rufing, and neurite extension [3336].

Src family kinases in tumor progression and metastasis

C-Yes is evidently unable to compensate for the


lack of c-Src in these processes. Regulation of
actin cytoskeleton dynamics is crucial to the
motility, invasiveness, and metastatic spread of
tumor cells, and thus the contributions of c-Yes to
colon carcinoma progression may differ from
those of c-Src. The data from Han et al. [32]
indicate that c-Src and c-Yes may be regulated
differently in colon cancer, and thus may make
unique contributions to colon cancer progression.
Interestingly, a third Src family kinase, Lck, is
also occasionally expressed in human colon
carcinoma cells [37]. This is of note due to the
fact that Lck expression is typically limited to cells
of hematopoietic origin. Lck expression in colon
cancer arises as a result of abnormal activation of
the type 1 Lck promoter, possibly due to the loss
of a transcriptional repressor [38]. The contribution of Lck to colon cancer progression, however,
has not been extensively studied and remains
poorly understood.

Src family kinases contribute directly to colon


tumor growth
The works cited above provide strong, but only
correlative evidence for a role for Src family
kinases in the progression of colon carcinoma
toward the malignant phenotype. Using cell line
models, several studies in recent years, however,
have provided more direct evidence that Src family
kinases play an important role in colon cancer progression. These studies involve ectopic expression
of c-Src or an activated variant or downregulation
of c-Src expression and/or kinase activity.

Ectopic expression of c-Src in colon cancer


In 1997, Irby et al. [39] transfected poorly
metastatic human colon tumor cell lines with a cSrc expression vector, resulting in clones that
expressed 410 fold higher levels of c-Src than
controls. These studies demonstrated that c-Src
overexpression did not affect cell proliferation in
vitro or metastatic spread of the tumor cells when
injected into nude mice, however, the growth of
primary tumors derived from these colon tumor
cells was signicantly enhanced in nude mice,

341

relative to wild type cells. The lack of effect on


metastatic spread of cancer cells in this study may
have been due to the fact that the overexpressed cSrc was not hyper-activated. Pories et al. [40]
demonstrated soft agar growth and increased
invasive potential upon overexpression of c-Src
in rat epithelial cell lines. Both of these studies
indicate that c-Src can contribute to different
aspects of tumor growth and/or progression
toward the metastatic phenotype.

Inhibition of Src family kinase activity in colon


tumor cells
Several methods have been used to downregulate
Src kinase activity, and these have provided
important contributions to the overall knowledge
of the role of Src family kinase activation in colon
cancer. In 1991, Garcia and colleagues [41]
demonstrated that treatment of human colon
tumor cell lines with the tyrosine kinase inhibitor
herbimycin A caused a reduction in c-Src kinase
activity, followed by subsequent reductions in cSrc protein levels and growth of colon tumor cells.
Herbimycin A treatment resulted in 40% growth
inhibition of colon tumor cell lines, as compared
with a 12% growth inhibition of a normal colon cell
line [41]. Kawai et al. [42] recently demonstrated
that herbimycin A treatment of colonic epithelial
cells expressing activated Src inhibited the proinvasive effects of TNF-a in these cells, indicating a
possible role for c-Src or SFK activation in the
progression of colon cells from inammatory
lesions such as ulcerative colitis to a cancerous
state. Interestingly, Novotny-Smith and Gallick
[43] demonstrated that prolonged TNF-a treatment
of TNF sensitive derivatives of the HT-29 colon
carcinoma cell line displayed reduced c-Src kinase
activity and cell proliferation in response to treatment with TNF-a. However, the reduction in cell
proliferation occurred prior to the decrease in c-Src
activity and thus was not likely caused by it.
Another strategy used to downregulate Src family
kinase activity is overexpression of the Src family
negative regulator Csk. Nakagawa et al. [44] utilized
an adenovirus expression system to overexpress Csk
in a mouse colon adenocarcinoma cell line. Cells
expressing Csk were able to form primary tumors in
mice but were decient in metastatic spread and in

342

Summy and Gallick

their ability to invade through Matrigel in vitro.


More recently, Boyer et al. [45] produced similar
results, expressing Csk or dominant negative
(kinase dead) Src in a rat carcinoma cell line. These
cells again were decient in their ability to form
metastases when implanted in mice, while still
retaining the ability to form primary tumors.
Overexpression of Csk or dominant negative Src
are useful tools for studying the collective role(s)
of Src family kinases in biologic systems. However,
because these strategies result in inhibition of
multiple Src family members (and often other
signaling molecules as well) they are unable to
provide insights with respect to contributions of
individual family members. To circumvent this
problem, Staley et al. [46] utilized a c-Src antisense
construct to effect exclusive downregulation of
c-Src in the colon cancer cell line HT-29. Clonal
transfectants obtained with the c-Src antisense
construct proliferated more slowly than parental
cells or cells transfected with a sense construct in
vitro. When injected into nude mice, the antisenseexpressing cells formed dramatically slower-growing tumors than did sense-expressing or wild type
HT-29 [46]. Not only do these data indicate the
importance of c-Src in the growth of human colon
tumor cells, but they reveal the importance of
specicity between Src family kinases, as c-Yes
protein expression levels and kinase activity were
not affected by the c-Src antisense construct.

Mechanisms of Src family kinase activation in


colon cancer
The studies discussed above indicate the importance of Src family kinase activation in the
progression of colon tumors; however, they do
not address the question of how Src family kinases
are activated in human colon cancer. The best
current data suggest that multiple mechanisms
contribute to Src activation in colon tumor cells.
Di Domenico and co-workers demonstrated that
both c-Src and c-Yes are activated upon estradiol
stimulation of Caco-2 colon carcinoma cells [47].
The method of activation was not determined;
however, the elevated tyrosine kinase activity
appeared to be important for downstream MAP
kinase activation and cell proliferation. Mao and
colleagues [48], in 1997, demonstrated that highly

metastatic colon cancer cell lines displayed elevated c-Src kinase activity that was further
enhanced upon stimulation with various mitogens.
Specically, stimulation of the EGF receptor,
Her2/Neu, and c-Met were found to induce c-Src
activation. EGFR was demonstrated to associate
with c-Src, indicating a potential SH2 displacement
mechanism for activation of c-Src in response to
these mitogens [48]. C-Src also associates similarly
with the protein tyrosine kinase receptor for
hepatocyte growth factor/scatter factor, c-Met.
We have demonstrated that Src activation resulting
from overexpression of c-Met contributes to the
tumorigenic and metastatic potential of colon
tumor cells (Herynk et al., submitted).
The regulation of Src family kinases by phosphorylation/dephosphorylation of the regulatory
tyrosine residue has been studied in some detail in
human colon cancer. In 1987, DeSeau et al. [49]
reported that despite no detectable differences in
tyrosine phosphatase activity between lysates of
normal colon cells and colon carcinoma cells,
treatment of cells with the phosphatase inhibitor
sodium vanadate resulted in increased c-Src kinase
activity from normal colon cells but not colon
tumor cells. These experiments indicate differences
in c-Src regulation between normal and cancerous
colon cells. Park and Cartwright [50] reported in
1995 that c-Src activity was elevated in colon
cancer cells above normal colonic epithelium and
was further increased 2- to 3-fold during mitosis of
colon carcinoma cells, indicating a potential role
for c-Src in progression through the cell cycle. This
increased activity corresponded with a decrease in
phosphorylation at the regulatory tyrosine residue.
Interestingly, in these cells, c-Yes activity and
protein levels decreased during mitosis, again
indicating differential regulation, and thus potentially differential functions, between c-Src and cYes in colon cancer [50]. Peng and Cartwright [51]
reported that Src and the tyrosine phosphatase
SH-PTP2 (Syp) were capable of co-association and
that this co-association resulted in phosphorylation of Syp by active Src and dephosphorylation of
Src Y527 by Syp, causing activation of both
enzymes. Zheng et al. [52] reported that Src
phosphorylation of PTP alpha at tyrosine 789
results in PTP alpha binding to the Src SH2
domain and dephosphorylation of Src at Y527
[52]. These studies indicate that c-Src may be

Src family kinases in tumor progression and metastasis

activated through both SH2 displacement and


Y527 dephosphorylation.
Phosphorylation of Y527, a critical aspect of
regulation of Src, is accomplished by Csk and its
family members. Cam and colleagues [53] demonstrated a mild reduction in Csk message levels,
protein levels, and kinase activity in colon
carcinoma tissues and cell lines, relative to normal
colon cells, and this correlated inversely with c-Src
protein levels and kinase activity, which were
elevated in the cancerous cells.
There has been one report of mutational
activation of c-Src in colon cancer. In 1999, Irby
et al. [14] reported a truncating mutation at residue
531 that resulted in an activated and transformation-competent c-Src variant in 12% of advanced
human colon cancers tested. However, this mutation has not been detected by other groups in
subsequent studies [5456]. Thus, while mutations
in Src that directly contribute to metastasis
provide a very important demonstration of the
role of Src in tumor progression, they do not
appear to represent a common mechanism of c-Src
activation in human colon cancer.
Finally, Rajala et al. [57] provided evidence that
c-Src protein levels and kinase activity may be
regulated in part by N-myristoyltransferase activity. These studies demonstrated that N-myristoyltransferase activity is higher in colonic epithelial
neoplasms than normal colonic tissue and that this
corresponds with increased levels of c-Src and
decreased expression of the N-myristoyltransferase
inhibitory protein (NIP71). Collectively, these data
suggest that, in most colon tumor cells, Src
activation is secondary to a variety of epigenetic
events (e.g. overexpression of growth factor
receptors, changes in integrin composition/function, changes in Csk regulation) that occur during
colon tumor progression. However, once activated, Src further contributes to the process.

Functions of Src family kinases in colon


cancer progression
Src family kinases in colon cancer motility,
invasiveness, and detachment
The above mentioned studies by Staley et al. [46]
and Irby et al. [14,39] indicate that c-Src activity is

343

important for proliferation and growth of primary


tumors derived from human colon cancer cells.
Aside from the role of c-Src in proliferation,
several recent studies indicate that Src contributes
to the motility and invasiveness of colon cancer
cells. It has long been known that c-Src plays an
important role in the dynamic regulation of the
actin cytoskeleton, and processes dependent on
this regulation, including cell motility, cell/matrix
adhesion, and cell/cell adhesion [8]. Brunton et al.
[58] demonstrated that during the adenoma to
carcinoma transition of colonic epithelia, levels of
EGF receptor and FAK were increased. Furthermore, EGF induced movement of the cells into the
basement membrane, coinciding with increased cSrc kinase activity, redistribution of c-Src to the
cell periphery, and enhanced phosphorylation of
FAK.
C-Src or Src family kinases may also contribute
to the invasiveness of colon carcinoma cells
through upregulated expression of matrix proteases that facilitate the degradation of the basement membrane necessary for tumor invasiveness.
Allgayer et al. [59] demonstrated that activated cSrc can induce transcription of the urokinase
plasminogen activator receptor, and transcription
of the gene for this receptor is blocked by
inhibitors of Src family kinases. Additionally,
Nakagawa et al. [44] demonstrated that overexpression of Csk in colon tumor cells results in
reduced expression of the matrix metalloprotease
MMP-2.
Another important event in the progression of
colon cancer to the metastatic phenotype is the
disruption of cadherin-mediated cell/cell contacts.
Studies in other systems have indicated that Src
family kinase activity will disrupt cell/cell junctions mediated by cadherins [8,60]. Irby and
Yeatman [61] recently demonstrated that expression of active c-Src in colon cancer cell lines
resulted in up to a 7-fold decrease in cell/cell
adhesion, accompanied by cadherin phosphorylation and loss of cadherin/catenin association.
Expression of dominant negative Src or incubation
with a Src family kinase inhibitor restored adhesion, demonstrating that the effect was specic for
activated Src family kinases. In a recent study,
Avizienyte et al. [62] went a step further and
attempted to determine whether it was the kinase
activity of activated c-Src or the enhanced avail-

344

Summy and Gallick

ability of the SH3 and SH2 domains that was


responsible for de-regulation of the E-cadherin
system in colon cancer cells. In this study, they
demonstrated that elevated c-Src activity caused
several of the components of adherens junctions to
re-distribute to integrin-based adhesion complexes, formed upon expression of active Src.
Additionally, they showed that while the noncatalytic domains were sufcient for formation of
integrin-adhesion complexes, they were not sufcient to induce disruption of E-cadherin containing cell/cell contacts [62].
For cells to metastasize, they must detach from
the primary tumor and surrounding cellular
matrix. This loss of attachment in normal cells
results in a form of programmed cell death termed
anoikis. Windham et al. [170] recently demonstrated that c-Src activation contributes to anoikis
resistance in colon carcinoma cells, through the
phosphatidylinositol 3-kinase (PI3K)/Akt pathway. In these studies, inhibiting Src family kinase
activity or downregulation of c-Src specically
through an antisense construct increased the
susceptibility of HT29 cells to anoikis. Conversely,
overexpression of activated c-Src in colon cancer
cells that express low levels of endogenous c-Src
enhanced the resistance of these cells to anoikis.
Increased c-Src expression/activity resulted in
increased Akt phosphorylation, but not that of
Erk 1/2.

decreased VEGF expression in HT29 colon


carcinoma cells, but the ability of hypoxia to
induce VEGF expression was decreased by over 50
fold in the c-Src antisense cells. Tumors formed by
these cells in nude mice showed a signicant
decrease in vascularity, in comparison to parental
cells, thus indicating that c-Src specically may
play an important role in the angiogenesis of colon
cancer cells in which it is activated. These data
suggest a biologic mechanism by which Src
activation might contribute to tumor progression,
as increased activity of Src would contribute to
angiogenesis in the growing tumor.
In a nal note on the role of c-Src in colon
cancer progression, Malek et al. [66] recently used
gene chip technology to compare the expression of
genes in human colon cancer cells to those induced
in Src-transformed rat broblasts. Several genes
were found to be transcriptionally activated in
both the Src-transformed broblasts and the colon
tumor cells, including cell cycle proteins, cytoskeletal proteins, and transcription factors. Technology such as this may help provide a more detailed
understanding of genes that are transcribed downstream of activated c-Src in colon carcinoma cells.
Such knowledge would certainly be useful in
improving the understanding of the roles of c-Src
and related proteins in the progression of colon
carcinoma.

Src family kinases in colon cancer angiogenesis

Breast carcinoma

The role of SFKs in colon cancer angiogenesis has


come under more intense investigation in recent
years. A critical study of Mukhadopathyay [63]
demonstrated that Src, but not its family members,
was critical to hypoxia-induced expression of
vascular endothelial growth factor (VEGF) in
broblasts. Thus, given the above studies demonstrating the importance of Src in tumorigenic
growth, and the slow rate of growth of tumor cells
resulting from Src reduction, the possibility that
Src may be exerting its effect, at least in part
through VEGF expression has been investigated.
Fleming et al. [64] demonstrated that levels of
vascular endothelial growth factor, an important
angiogenic factor, varied directly with c-Src levels.
Ellis et al. [65] demonstrated that not only did
antisense-mediated c-Src downregulation result in

Although the functions of Src family kinases in


breast cancer have not been studied to the extent
that they have in colon cancer, considerable data
support a role for SFKs in the progression of this
cancer as well. In 1983, Jacobs and Rubsamen [67]
demonstrated elevated c-Src kinase activity in
breast carcinoma tissue, in comparison to normal
breast epithelium. In 1986, Rosen et al. [18]
reported similar ndings. In 1989, Lehrer and
colleagues [68] attempted to determine if elevated
c-Src kinase activity was associated with any
clinical parameters. Their studies indicated that
tumors expressing the progesterone receptor generally displayed higher c-Src kinase activity than
those that did not. Koster et al. [69] demonstrated
that 2530% of breast cancer specimens studied
displayed signicant expression of at least one

Src family kinases in tumor progression and metastasis

proto-oncogene, including c-src, erbB, raf1, lck,


and H-ras. In the rst comprehensive study of
human breast tumors, Ottenhoff-Kalff et al. [70]
reported in a study of 72 breast cancer specimens,
that all tumor samples displayed elevated tyrosine
kinase activity relative to normal controls and that
70% of this tyrosine kinase activity could be
attributed to c-Src. In 1996, Verbeek and coworkers [71] utilized immunohistochemical staining, along with in vitro kinase assays and western
blots, to demonstrate that c-Src protein expression
and kinase activity were elevated in breast cancer
tissue, relative to normal breast tissue. Similar
results were obtained by Reissig and colleagues
[72]. In both studies, c-Src protein levels and
kinase activity were found to be elevated relative
to non-cancerous tissue.

Mechanisms of Src family kinase activation in


breast cancer
As with colon cancer, multiple mechanisms lead to
SFK activation in breast carcinomas. Breast
cancer cells frequently overexpress members of
the EGFR family of receptor tyrosine kinases,
including EGFR and Her2/Neu [73]. As SFKs are
known to be activated downstream of receptor
tyrosine kinases via displacement of the tail
phosphotyrosine from the SH2 domain with
higher afnity phosphotyrosine residues on the
receptor [7478], this may represent an important
mechanism of SFK activation in breast cancer
cells. Luttrell et al. [79] demonstrated that the cSrc SH2 domain associated with activated EGFR
and Neu from breast cancer cell lines. Additionally, endogenous c-Src was shown to associate
with the activated EGFR. In 1994, Oude Weernink et al. [80] demonstrated EGF-dependent
enhanced c-Src kinase activity from cells overexpressing EGFR. Meanwhile, Muthuswamy and
co-workers [81,82] found evidence for c-Src
activation downstream of Neu in a mouse model
system. In this system, transgenic mice develop
mammary tumors as a result of Neu expression
under the control of the mouse mammary tumor
virus promoter/enhancer. C-Src activity was
shown to be 6 to 8 fold higher in tumors than
the surrounding mammary epithelium, and c-Src

345

formed a stable complex with Neu, indicating that


c-Src activation was likely due to an SH2
displacement mechanism [82]. Muthuswamy and
Muller [83] later demonstrated that c-Src association with Neu was dependent on tyrosine phosphorylation of Neu, and was likely dependent on
the c-Src SH2 domain. Interestingly, in their
system, c-Src did not associate with EGFR, and
EGF stimulation of cell lines overexpressing
EGFR resulted in transphosphorylation of Neu
and co-association of Neu and c-Src. This same
group of investigators later reported that c-Yes
was also activated in Neu-induced mouse mammary tumors and this activation correlated with
the ability of c-Yes to associate with Neu in an
SH2-dependent fashion [84]. Biscardi et al. [85]
found that c-Src and EGFR co-associate in an
EGF-dependent fashion in breast cancer cell lines
and tumor samples which overexpress both
proteins. Discrepancies in the ndings on the coassociation between c-Src and EGFR may be due
to differences in the model systems used in the
various studies. It is also possible that c-Src may
interact with the EGFR indirectly through an
intermediary protein. In support of this latter
possibility, Li et al. [86] demonstrated that the
mucin-like transmembrane glycoprotein MUC1
constitutively associates with EGFR in breast
carcinoma cells. Activation of EGFR in these cells
results in phosphorylation of MUC1 on tyrosine
and subsequent SH2-dependent binding of c-Src to
MUC1.
In recent years, several studies have indicated
that SFKs may be activated in breast cancer via
tyrosine phosphatase-dependent dephosphorylation of the regulatory tyrosine. Egan et al. [87]
observed that c-Src kinase activity from breast
cancer cells was elevated 5.6 fold above that
detected in normal control cells and that this
elevated activity could be suppressed when the
cells were grown in the presence of the phosphatase inhibitor sodium vanadate. Furthermore, a
tyrosine phosphatase activity capable of dephosphorylating the regulatory tyrosine was detected
in membrane fractions from breast tumor cells.
This phosphatase was later identied as proteintyrosine phosphatase 1 B (PTP1B). In this study,
Bjorge and colleagues [88] demonstrated further
that overexpression of PTP1B in 293 cells resulted
in a 2-fold increase in c-Src activity. It should be

346

Summy and Gallick

noted again that SH2 displacement and dephosphorylation of the regulatory tyrosine are not
mutually exclusive and both may contribute to
SFK activation in breast cancer cells.

Contributions of SFKs to breast cancer


progression
As with colon cancer, activation of SFKs in breast
cancer does not guarantee a signicant contribution to the biological progression of these tumors.
However, substantial data exist to indicate that
this is indeed the case. The transgenic mouse
model has been important in determining the
importance of SFKs in mammary tumorigenesis.
Webster and colleagues utilized transgenic mice to
express an activated c-Src specically in mouse
mammary glands, under the transcriptional control of the mouse mammary tumor virus long
terminal repeat [89]. Further, female mice expressing active c-Src frequently developed epithelial
hyperplasias, which occasionally progressed to full
neoplasias [89]. In an alternative mouse model,
mice harboring the polyomavirus (PyV) middle T
antigen under the control of a mammary-specic
promoter invariably develop mammary tumors.
However, when mice lacking the c-src gene express
PyV middle T, mammary tumor formation occurs
less frequently and at a delayed rate [90]. These
results indicate an important role for c-Src in the
formation of mammary tumors in this system.
Interestingly, mice lacking the c-yes gene develop
PyV middle T-induced mammary tumors at a
normal rate; however, middle T is decient in
transformation of endothelial cells in the absence
of c-Yes [90]. These results indicate that c-Yes may
not be important for mammary tumorigenesis, at
least in this model system, and further illustrate
the lack of absolute functional redundancy between
c-Src and c-Yes, despite their signicant homology.

Src family kinases in breast cancer cell growth and


survival
While the transgenic mouse model is useful,
extrapolation of results obtained in this system
to human breast cancer can be problematic.
Several studies, however, have focused on the

importance of c-Src and SFKs in the growth and


survival of human breast cancer cells. As mentioned above, breast tumors with elevated c-Src
activity frequently express the progesterone receptor, raising the possibility that SFK may contribute to hormone-dependent cell growth
signaling [68]. In support of this possibility,
Arnold et al. [91] demonstrated that the human
estrogen receptor was phosphorylated by c-Src in
vitro. This group determined also that tyrosine
phosphorylation of the human estrogen receptor
was necessary for dimerization of the receptor and
binding to the estrogen receptor response element
[92]. Migliaccio et al. [93] demonstrated that c-Src
is activated downstream of the estradiol receptor
in a hormone-dependent fashion, leading to
subsequent activation of the mitogen activated
protein kinase (MAPK) pathway. This same group
later demonstrated that progestins acting through
the progesterone receptor can induce activation of
c-Src, and subsequent MAPK pathway activation,
through cross-talk with the estrogen receptor [94].
However, Boonyaratanakornkit et al. [95] demonstrated that the progesterone receptor was capable
of directly activating SFKs, through displacement
of the SH3 domain/linker intramolecular interaction, in response to progestin. Thus, while SFK
activation is evidently necessary for progesterone
receptor cell signaling, the mechanism of SFK
activation downstream of hormone stimulation is
unclear and may vary from system to system.
In estradiol-stimulated MCF-7 cells, SFK activity is necessary for activation of PI3K and Akt,
leading to subsequent cyclin D1 transcription and
cell cycle progression [96]. Furthermore, Tsai and
colleagues demonstrated that estrogen receptornegative breast cancer cell lines were still capable
of inducing Akt activation upon estrogen stimulation in a SFK-dependent fashion [97]. Collectively,
these results suggest that c-Src or SFKs may be
important for breast cancer cell growth and
survival stimulated by multiple hormones.
Enhanced expression of c-Src in cells expressing
elevated levels of EGFR resulted in increased
DNA synthesis, soft agar growth and tumor
formation in nude mice [98]. Shefeld and
colleagues demonstrated that activation of c-Src
in mammary epithelial cells overexpressing ErbB2
and further showed that inhibition of c-Src activity
reduced the ability of these cells to grow in soft

Src family kinases in tumor progression and metastasis

agar [99]. In 1998, Amundadottir and Leder [100]


reported that inhibition of SFK activity with the
Src family inhibitor PP1 disrupted anchorageindependent growth of a TGF-a induced mammary tumor cell line. c-Src is also activated
downstream of broblast growth factor-2 (FGF2) signaling in the human breast cancer cell line
MCF-7 [101]. In this system, c-Src activation led to
the tyrosine phosphorylation of the cell cycle
mediator cyclin D2. Recently Belsches-Jablonski
et al. [102], demonstrated overexpression and coassociation of Her2/Neu and c-Src in human
breast cancer cell lines and tumor samples. In
these studies, heregulin treatment stimulated
anchorage-independent growth, and inhibition of
SFK activity with the inhibitor PP1 resulted in
reduced anchorage-dependent and anchorageindependent growth and triggered apoptosis,
indicating a role for SFKs in both growth and
survival of breast carcinoma cells. Disruption of
SFK activity in several human cancer cell lines,
including those of breast cancer origin, induced a
mitotic block that occurred after chromosome
condensation and before spindle assembly [103].
A cooperative effect of activated c-Src and Stat3
in the transcriptional activation of the gene for
hepatocyte growth factor (HGF), also known as
scatter factor has been demonstrated [104]. HGF/
c-Met signaling facilitates cell growth, motility,
and survival [105], and thus represents another
system wherein c-Src activation may contribute to
the metastatic progression of human breast cancer
cells. In human breast cancer cell lines expressing
elevated EGFR and c-Src kinase activities, Stat3 is
constitutively activated in a SFK and JAKdependent fashion [106]. SFK inhibition resulted
in a reduction of Stat3 DNA binding, growth
inhibition, and initiation of apoptosis. Olayioye et
al. [107] recently demonstrated that SFK activity is
essential for EGF and neuregulin induced MAPK
activation in multiple breast carcinoma cell lines.
Bougeret and colleagues [108] reported that the
Csk homologous kinase Chk, when transfected
into MCF-7 breast carcinoma cells, inhibits heregulin induced activation of c-Src and Lyn and
delays mitotic entry. It was subsequently shown
that Chk expression inhibits migration of T47D
breast cancer cells, presumably due again to
inactivation of SFKs [109]. Thus, Src may be a
critical regulator of a variety of protein tyrosine

347

kinase receptors aberrantly expressed and/or


activated in breast cancer.

Src family kinases in breast cancer angiogenesis


As with colon cancer, c-Src is important in the
induction of VEGF transcription in breast cancer
cells. Expression of the antisense c-Src vector
described by Pal and co-workers resulted in
decreased VEGF transcription in breast cancer
cells. Interestingly, VEGF expression and c-Src
activation were inhibited under hypoxic conditions
by the tumor suppressor p53. Taken together,
these results suggest an important role for c-Src in
a variety of signaling pathways involved in breast
cancer cell growth, survival, and angiogenesis.

Src family kinases in detachment, motility, and


invasiveness of breast carcinoma
c-Src is necessary for hyaluronan mediated motility (RHAMM) motility of cells [36]. It has been
demonstrated previously that RHAMM is
required for breast cancer motility, and thus cSrc may likewise be essential for motility in breast
cancer cells [111]. In a study of RAFTK function
in breast cancer cells, [112] Zrihan-Licht and
colleagues demonstrated that after HRG stimulation, RAFTK associated with p190 RhoGAP,
RASGAP, and ErbB-2, and is required for
phosphorylation of p190 RhoGAP by Src. Association of RAFTK with ErbB-2 was abolished by
mutation of the Src binding site, as was the
stimulation of invasion normally detected upon
expression of wild type RAFTK in breast cancer
cells [112]. These results indicate that c-Src, in
conjunction with RAFTK may play a pivotal role
in the invasion of breast cancer cells. Src family
kinases may also contribute to motility through
shedding of the L1 adhesion molecule [113]. These
studies demonstrated that upon pervanadate
treatment or EDTA-mediated cell detachment,
the SFK Fyn was activated, and the cytoplasmic
tail of L1 was subsequently tyrosine phosphorylated. Incubation with the SFK inhibitor PP2
blocked cell rounding and detachment from the
substrate. The authors proposed that, as soluble
L1 binds to the proteoglycan neurocan, L1

348

Summy and Gallick

shedding may support integrin-mediated cell


migration and adhesion [113]. As discussed above,
Hung and Elliott [104] demonstrated that continuous activation of c-Src, in conjunction with Stat3,
can result in elevated HGF expression, and
subsequent activation of autocrine HGF/c-Met
signaling. In addition to promoting tumor cell
growth, this signaling pathway may contribute to
the motility and invasiveness of breast cancer cells.
In support of this idea, Rahimi et al. [114]
demonstrated that c-Src is activated in response
to HGF stimulation of breast carcinoma cells and
breast epithelial cells and that this activation
corresponds with co-association between c-Met
and c-Src. Expression of dominant negative Src in
these cells disrupted HGF-induced motility and
soft agar growth but did not affect cell proliferation on a plastic substratum. SFKs may also play a
role in downregulation of E-cadherin mediated cell
aggregation in breast cancer cells, as PP2 stimulated aggregation of breast cancer cells in their
studies [115]. Finally, it was recently demonstrated
that treatment of breast cancer cells with the SFK
inhibitor PP1 or expression of a dominant negative
Src vector signicantly reduced CD99-induced cell
motility [116]. Taken together these data indicate
that not only do SFKs likely contribute to the
growth and survival of breast cancer cells, but also
to motility, invasiveness, and thus metastatic
potential.

c-Yes activation in human melanoma


In comparison to breast and colon cancer, much
less is known about the roles of SFKs in other
human cancers; however, the results of several
studies suggest that they may be involved in a
variety of human epithelial and non-epithelial
cancers. Human melanoma is one of the few
cancers in which c-Yes has been more strongly
implicated than c-Src. Increased expression and
kinase activity of c-Yes has been demonstrated in
human melanoma cell lines, relative to normal
melanocytes [117]. No differences were detected in
the expression and activation of c-Src between the
melanoma cells and melanocytes. Later, Marchetti
et al. [118] showed that c-Yes tyrosine kinase
activity was further elevated in brain metastatic
melanoma cell lines. Furthermore, they demon-

strated that c-Yes was activated upon stimulation


with the neurotrophin nerve growth factor (NGF),
while c-Src activity was unaffected. Further
supporting the potential link between c-Yes
activity and metastatic progression, the degree of
c-Yes activation in response to NGF stimulation
correlated with the invasiveness of the melanoma
cells [118]. These studies again indicate specicity
in signaling between c-Src and c-Yes in human
cancers and suggest that c-Yes may make a more
signicant contribution than c-Src to the progression of human melanoma.

Src family kinases in pancreatic cancer


In 1996, Visser et al. [119] utilized a rat experimental exocrine pancreatic carcinogenesis model
system to demonstrate that c-Src protein expression and kinase activity were elevated in azaserine
treated rat pancreas, with respect to untreated
controls. c-Src overexpression was observed in 13/
13 pancreatic carcinoma cell lines, in comparison
to normal pancreatic specimens and elevated c-Src
kinase activity in 14/17 pancreatic carcinoma cell
lines, when compared to normal pancreatic cells
[120]. In these studies, c-Src kinase activity did not
correlate with protein expression levels or expression of Csk. MacMillan-Crow et al. [121] recently
demonstrated that c-Src was subject to tyrosine
nitration in pancreatic carcinoma cells, and that
this correlated with a greater than two fold
increase in c-Src kinase activity and association
between c-Src and cortactin. Tyrosine nitration
may occur in the presence of oxidative stress as a
result of increased levels of the reactive nitrogen
species peroxynitrite [121]. This study indicates
that tyrosine nitration may contribute to c-Src
activation, and subsequent downstream growth
and survival signals in the face of oxidative stress.
Activated c-Src expression in pancreatic carcinoma cells results in elevated expression of the
insulin-like growth factor 1 (IGF-1) receptor [122].
Upregulation of the IGF-1 receptor in response to
Src may occur as a result of Akt activation
downstream of Src [123]. Elevated expression of
the IGF-1 receptor leads to increased IGF-1dependent cell proliferation, thus indicating
another potential mechanism by which c-Src may
contribute to pancreatic cell growth and progres-

Src family kinases in tumor progression and metastasis

sion. As with breast and colon cancer, c-Src


activation not only contributes to pancreatic
cancer growth but also to invasion and metastasis.
The growth of pancreatic carcinoma cell lines with
high metastatic potential on type III collagen
results in reduced E-cadherin expression,
decreased cellcell adhesion, and enhanced migration [124]. These results were also obtained upon
overexpression of activated c-Src. E-cadherin
downregulation in response to collagen could be
suppressed by treatment with the SFK inhibitors
PP1 and herbimycin A [124]. These studies indicate
a potentially important role for c-Src or SFKs in
the growth and progression of human pancreatic
cancer.

Src family kinases in ovarian cancer


As with colon and breast cancer, expression of an
antisense c-src vector in human ovarian cancer
cells indicates that c-Src may play a specic and
vital role in human ovarian cancer. Antisense c-src
vector expression in a human ovarian cancer cell
line resulted in decreased anchorage-independent
growth, decreased tumor growth in a mouse
model, and decreased VEGF expression and
tumor vascularization [125]. These results indicate
that c-Src may be important for several different
aspects of ovarian cancer growth and progression.
c-Src activity appears important for Shc phosphorylation and Erk1/2 activation downstream of
CXCR-1/2 receptor stimulation [126]. Finally, the
importance of Src was demonstrated in hyaluronic
acid-dependent ovarian cell tumor migration, as
overexpression of active Src stimulates HAdependent tumor cell migration, while expression
of dominant negative Src abrogates the tumorigenic phenotype [127].

349

(EMT) of bladder cancer cells was demonstrated


[129]. Overexpression of c-Src results in EMT or
sensitization to growth factor-mediated EMT.
Additionally, overexpression of dominant negative
Src inhibits growth factor induced cell scattering
that occurs concomitantly with EMT [129]. More
recently, utilizing human bladder carcinoma cell
lines, SFK inhibition was demonstrated to block
EMT [130], thus suggesting a potentially important role for SFKs in the progression of bladder
cancer cells to an invasive phenotype.

Src family kinases in gastric cancer


Takekura and colleagues [131], in 1990, demonstrated that c-Src kinase activity was elevated in 8
of 16 gastric carcinoma tissues, in comparison to
normal mucosa samples. The levels of kinase
activity did not necessarily correlate with levels
of c-Src protein expression.

Src family kinases in head, neck, and esophageal


cancers
In 1992, Jankowski et al. [132] detected c-Src
expression in Barretts esophagus (BE) and esophageal adenocarcinoma, but not in normal
mucosa. c-Src kinase activity was elevated three
to 4 fold in BE and 6 fold higher in esophageal
adenocarcinoma than normal control tissues [133].
Dephosphorylation of c-Src at the regulatory
tyrosine was detected in BE tissue. c-Src is
overexpressed in hyperproliferating regions of
head and neck squamous cell carcinoma, dysplastic epithelium, benign papillomas, and inamed
normal tissue [134].

Src family kinases in lung cancer


Src family kinases in bladder cancer
In 1992 Fanning and colleagues [128] demonstrated c-Src kinase activity was elevated in a panel
of human bladder carcinomas over normal tissue
and that increased kinase activity was associated
primarily with low grade bladder lesions. In a rat
bladder carcinoma model the importance of
c-Src in the epithelium-to-mesenchyme transition

In 1990, Kiefer et al. [135] demonstrated expression of v-src-related mRNA in a panel of nonsmall-cell lung cancer (NSCLC) lines, with elevated expression detected in one adenocarcinoma
cell line. In a panel of lung cancers Src expression
was found in 20 of 33 small cell lung cancer and
NSCLC samples but not in histologically normal
lung tissue [136]. In the NSCLC samples, 6080%

350

Summy and Gallick

of adenocarcinomas and bronchiolo-alveolar cancers and 50% of squamous cell carcinomas


displayed c-Src protein expression. C-Src expression was detected more frequently in poorly
differentiated squamous cell carcinomas than in
moderately or well differentiated samples [136].
Consistent with these observations, c-Src was
found to be activated in NSCLC cell lines [137].
While relatively little is known of the contributions of SFKs to lung cancer growth and progression, in comparison to other cancers of epithelial
origin, recent studies have begun to shed light on
this issue. Lck and c-Yes kinase activities were
shown to be elevated in a small cell lung cancer cell
line downstream of stem cell factor (SCF)mediated Kit stimulation [138]. Inhibition of
SFK activity with PP1 resulted in a complete
block of SCF-mediated cell growth [138]. It was
later demonstrated that Lck was required for SCFmediated MAPK activation; however, this pathway was dispensable for cell growth in the absence
of Rb [139]. Finally, evidence of a role for SFKs in
hypoxic growth and angiogenesis of human lung
cancer cells has been described [140]. Under
hypoxic conditions, the lung adenocarcinoma cell
line A549 resulted in transcriptional upregulation
of the endothelial PAS domain protein-1 (EPAS-1)
in a SFK-dependent fashion. EPAS-1, in turn,
upregulates its own expression and that of VEGF
[140]. These data indicate that SFKs may be
involved in the regulation of signaling pathways
that govern multiple aspects of lung cancer
progression.

Src family kinases in brain and neuronal tumors


Given the high expression levels of SFKs, including c-Src, Fyn, and c-Yes, in cells of neuronal
origin, it might be expected that they play an
important role in the growth and progression of
brain and neuronal tumors. While not as extensively studied as colon and breast cancer, the
expression and activation of c-Src in brain cancers
has been examined in a number of studies since
the mid-1980s. In 1985, Takenaka and colleagues
[141] demonstrated c-Src protein expression in
human astrocytomas but not normal brain tissues.
A 20 to 40 fold elevation in c-Src kinase activity
was observed in human neuroblastoma cell lines,

in comparison to glioblastomas or normal human


broblasts [142]. Later work demonstrated that
the abundance of c-Src in neuroendocrine tumors
varied with the neuronal differentiation state of
the cells and that the specic kinase activity of
c-Src from neuroblastoma cells, when normalized
against protein levels, was similar to that from
human glioblastoma cells [143]. In other studies,
c-Src kinase activity was elevated in neuroendocrine-derived neuroblastoma and small-cell lung
carcinoma cells, relative to non-neurocrine cells
[144], and c-Src protein levels and kinase activity
correlate with the differentiation state of neuroendocrine-derived tumors [145]. c-Src is expressed in
childhood neuroblastoma and retinoblastoma,
with the neuronal c-Src splice variant (c-SrcN)
being more highly expressed than c-Src in tumors
associated with good clinical prognosis [146].
Expression of v-Src under the control of the glial
brillary acidic protein (GFAP) regulatory elements in transgenic mice resulted in the formation
of astrocytomas in the brain and spinal cord [147].
As the tumors progressed, they became highly
proliferative and demonstrated greater malignancy. The tumors were also highly vascularized
in areas, resembling human glioblastomas [147].
VEGF expression occurs early in development of
these tumors preceding vascularization, and likely
contributes to the progression of these Srcinduced tumors [148]. While data on how c-Src
or SFKs may contribute to the growth and
progression of brain and neuronal tumors is still
relatively sparse, a recent study indicates that
activated and co-associated focal adhesion kinase
(FAK) and c-Src may contribute to the proliferation of anaplastic astrocytoma cells through
ERK2 activation and induction of cyclin D and
E expression [149].

Src family kinases in leukemias, lymphomas, and


myelomas
Given that the majority of SFKs are expressed
predominantly in cells of the hematopoietic lineage
and that SFKs are commonly involved in the
signaling pathways that regulate cell growth and
division, it might be reasonable to assume that
SFKs may be involved in the growth and
progression of cancers arising from cells of

Src family kinases in tumor progression and metastasis

hematopoietic origin. Interestingly, however, there


is relatively little evidence to denitively implicate
SFK activity in these processes, possibly due to the
absolute requirement for SFKs in the function of
these cells. The SFK Lyn is predominantly
expressed in B lymphocytes and monocytes/
macrophages, and several reports in the literature
are suggestive of a role for Lyn in cancers derived
from these cells. Lyn kinase activity, as opposed to
those of Fyn, Yes, or Hck, was specically
activated in myeloid leukemia cell lines in response
to IL-3, an important promoter of growth and
survival [150]. These results indicate that Lyn may
be involved in IL-3 dependent signal transduction
in human myeloid leukemia. Lyn may also be
involved in IL-6 mediated cell proliferation, as
activation of Lyn downstream of IL-6 stimulation
was demonstrated in multiple myeloma cells [151].
More recently, expression of a Lyn antisense
oligonucleotide or incubation with a SFK-specic
tyrosine kinase inhibitor reduced IL-6-dependent
proliferation of CD45 myeloma cells [152]. Lyn
also appears to be preferentially expressed in
human malignant lymphomas of B cell origin
[153]. When complexed with CD19 at the plasma
membrane of B cell lymphomas, Lyn may be an
important regulator of cell survival [154]. Lyn,
Raf-1, and MAP kinases are activated upon crosslinking of the intercellular adhesion molecule 1
(ICAM-1) in a B cell lymphoma line [155].
Katagiri et al. [156] demonstrated that Lyn, in
conjunction with fellow SFK Fgr, may be important for the prevention of apoptosis during
granulocytic differentiation of a promyelocytic
leukemia cell line, as inhibitors of tyrosine kinase
activity or antisense oligonucleotides induced
premature apoptotic cell death during retinoic
acid-induced differentiation. Expression of a Lyn
antisense oligonucleotide resulted in reduced proliferation of a myeloid leukemia cell line in
response to granulocyte/macrophage colony stimulating factor (GM-CSF) [157]. These results
were mirrored by incubation with the SFK
inhibitor PD166285, which also inhibited the
growth of leukemic cell lines. Expression of a
dominant negative Lyn (Y397F) in erythroleukemic cells had a negative effect on the development
of erythroleukemias in vivo, resulting in a reduced
mortality rate and a lengthened latency period
[158]. Finally, Ptasznik et al. [159] uncovered

351

evidence that Lyn may contribute to BCR/ABLdependent cell motility. Lyn is normally activated
downstream of stromal-derived factor 1 (SDF-1)
stimulation of its receptor CXCR4, as part of a
signaling cascade regulating the migration of
progenitor cells [159]. In leukemic cells expressing
Bcr/Abl, Lyn is constitutively active as a result of
association with Bcr/Abl, and is not responsive to
SDF-1 stimulation. Inhibition of Lyn results in
disruption of Bcr/Abl induced cell motility [159].
Importantly, overexpression of Lyn has been
recently observed to be one mechanism by which
Bcr/Abl positive CML cells become resistant to the
inhibitor STI571 (Donato et al., [171]). Hck also
associates with Bcr/Abl and expression of dominant negative Hck resulted in suppression of
cytokine-independent growth of Bcr/Abl expressing myeloid leukemia cells [160]. In total, the
reports summarized here suggest that Lyn may be
involved in the growth, survival and motility of
multiple human cancers of hematopoietic cell
origin.
The Lck tyrosine kinase is expressed primarily
in T lymphocytes, where it is thought to play an
important role in T cell hematopoiesis, proliferation, and receptor signaling [161164]. However,
Lck expression has been detected in other
leukocytes as well. Expression of lck mRNA has
been observed in B cell chronic lymphocytic
leukemia in 1991 [165]. Lck is associated with
the CD19 receptor in B lineage acute lymphoblastic leukemia (ALL) cells and is activated upon
engagement of the CD19 receptor, suggesting a
potential role for Lck in regulation of apoptosis in
B-lineage ALL cells [166]. Lck may contribute to
the induction of cell proliferation by human T cell
leukemia virus type 1 (HTLV-1) [167]. In these
studies the viral oncoprotein p40tax-1 alone was
unable to stimulate cell proliferation in the
absence of cytokines. However, P40tax-1was able
to induce cytokine-independent proliferation in
the presence of constitutively activated Lck [167].
Lck is expressed in B cell chronic lymphocyte
leukemia (CLL) and CD5 B-1cells, a self-renewing subpopulation of B cells that may represent
the normal precursors of CLL cells [168]. Thus,
despite its predominant expression in T lymphocytes, Lck may be involved in the growth and
survival of cancers originating from multiple
blood cell lineages.

352

Summy and Gallick

Conclusion
In summary, there is a wealth of data available
that indicate the importance of SFKs in the
growth and progression of human cancers. While
much of that work has focused on the role of cSrc, and to a lesser extent c-Yes, in breast and
colon cancers, SFKs have been implicated in many
different epithelial and non-epithelial cancers.
Their contributions to the biology of these cancers
are likely due to both elevated kinase activity and
increased availability of the functional domains
for intermolecular interactions.

Key unanswered questions


A key issue to be resolved in the future will be the
viability of SFK targeting in anti-tumor therapy.
With the remarkable success of the relatively
selective protein tyrosine kinase inhibitor STI 571
in treatment of Chronic Myelogenous Leukemia
patients, interest in developing selective inhibitors
of other protein tyrosine kinases continues to
grow. Concerted efforts are being made to develop
Src inhibitors, and Metcalf et al. [169] estimate
that Src inhibitors will reach the clinic within two
years. Due to their importance to several different
aspects of cancer progression, including cell
growth, survival, motility, invasion, and angiogenesis, SFKs may represent ideal candidates for
targeted anti-cancer therapeutics. Further, due to
the normal role of Src in bone resorption, Src
inhibitors may hold great promise for inhibiting
progression of tumors with a propensity to
metastasize to the bone, such as prostate and
breast cancer, as well as multiple myeloma.
However, the best use of such inhibitors still
requires a basic understanding of the role of Src
activation in tumor progression. While considerable progress has been made, much remains to be
learned about a class of enzymes that play such
key intermediary roles in signal transduction.

Acknowledgements
This work was supported by NIH 2 R01 CA65527,
NIH 1 U54 CA090810, the Gillson Longenbaugh

Foundation, and the Lockton Fund for Pancreatic


Research.

References
1. Rous P: A sarcoma of the fowl transmissible by an agent
separable from the tumor cells. J Exp Med 13: 397411,
1911
2. Brugge JS, Erikson RL: Identication of a transformation-specic antigen induced by an avian sarcoma virus.
Nature 269: 346348, 1977
3. Stehelin D, Varmus HE, Bishop JM, Vogt PK: DNA
related to the transforming gene(s) of avian sarcoma
viruses is present in normal avian DNA. Nature 260: 170
173, 1976
4. Hunter T, Sefton BM: Transforming gene product of
Rous sarcoma virus phosphorylates tyrosine. Proc Natl
Acad Sci USA 77: 13111315, 1980
5. Thomas SM, Brugge JS: Cellular functions regulated by
Src family kinases. Annu Rev Cell Dev Biol 13: 513609,
1997
6. Martin GS: The hunting of the Src. Nat Rev Mol Cell Biol
2: 467475, 2001
7. Toyoshima K, Yamamoto T, Kawai S, Yoshida M: Viral
oncogenes, v-yes and v-erbB, and their cellular counterparts. Adv Virus Res 32: 97127, 1987
8. Frame MC: Src in cancer: Deregulation and consequences
for cell behavior. Biochim Biophys Acta 1602: 114130,
2002
9. Sicheri F, Moare I, Kuriyan J: Crystal structure of the
Src family tyrosine kinase Hck. Nature 385: 602609, 1997
10. Xu W, Harrison SC, Eck MJ: Three-dimensional structure
of the tyrosine kinase c-Src. Nature 385: 595602, 1997
11. Kmiecik TE, Shalloway D: Activation and suppression of
pp60c-src transforming ability by mutation of its primary
sites of tyrosine phosphorylation. Cell 49: 6573, 1987
12. Piwnica-Worms H, Saunders KB, Roberts TM, Smith
AE, Cheng SH: Tyrosine phosphorylation regulates the
biochemical and biological properties of pp60c-src. Cell
49: 7582, 1987
13. Cartwright CA, Eckhart W, Simon S, Kaplan PL: Cell
transformation by pp60c-src mutated in the carboxyterminal regulatory domain. Cell 49: 8391, 1987
14. Irby RB, Mao W, Coppola D, Kang J, Loubeau JM,
Trudeau W, Karl R, Fujita DJ, Jove R, Yeatman TJ:
Activating SRC mutation in a subset of advanced human
colon cancers. Nat Genet 21: 187190, 1999
15. Irby RB, Yeatman TJ: Role of Src expression and
activation in human cancer. Oncogene 19: 56365642,
2000
16. Brown MT, Cooper JA: Regulation, substrates and
functions of src. Biochim Biophys Acta 1287: 121149,
1996
17. Fidler IJ: The biology of cancer metastasis or, you cannot
x it if you do not know how it works. Bioessays 13: 551
554, 1991

Src family kinases in tumor progression and metastasis


18. Rosen N, Bolen JB, Schwartz AM, Cohen P, DeSeau V,
Israel MA: Analysis of pp60c-src protein kinase activity in
human tumor cell lines and tissues. J Biol Chem 261:
1375413759, 1986
19. Bolen JB, Veillette A, Schwartz AM, Deseau V, Rosen N:
Analysis of pp60c-src in human colon carcinoma and
normal human colon mucosal cells. Oncogene Res 1: 149
168, 1987
20. Cartwright CA, Kamps MP, Meisler AI, Pipas JM,
Eckhart W: pp60c-src activation in human colon carcinoma. J Clin Invest 83: 20252033, 1989
21. Cartwright CA, Meisler AI, Eckhart W: Activation of the
pp60c-src protein kinase is an early event in colonic
carcinogenesis. Proc Natl Acad Sci USA 87: 558562,
1990
22. Cartwright CA, Coad CA, Egbert BM: Elevated c-Src
tyrosine kinase activity in premalignant epithelia of
ulcerative colitis. J Clin Invest 93: 509515, 1994
23. Weber TK, Steele G, Summerhayes IC: Differential
pp60c-src activity in well and poorly differentiated human
colon carcinomas and cell lines. J Clin Invest 90: 815821,
1992
24. Talamonti MS, Roh MS, Curley SA, Gallick GE: Increase
in activity and level of pp60c-src in progressive stages of
human colorectal cancer. J Clin Invest 91: 5360, 1993
25. Termuhlen PM, Curley SA, Talamonti MS, Saboorian
MH, Gallick GE: Site-specic differences in pp60c-src
activity in human colorectal metastases. J Surg Res 54:
293298, 1993
26. Sakai T, Kawakatsu H, Fujita M, Yano J, Owada MK:
An epitope localized in c-Src negative regulatory domain
is a potential marker in early stage of colonic neoplasms.
Lab Invest 78: 219225, 1998
27. Iravani S, Mao W, Fu L, Karl R, Yeatman T, Jove R,
Coppola D: Elevated c-Src protein expression is an
early event in colonic neoplasia. Lab Invest 78: 365371,
1998
28. Aligayer H, Boyd DD, Heiss MM, Abdalla EK, Curley
SA, Gallick GE: Activation of Src kinase in primary
colorectal carcinoma: An indicator of poor clinical
prognosis. Cancer 94: 344351, 2002
29. Park J, Meisler AI, Cartwright CA: c-Yes tyrosine kinase
activity in human colon carcinoma. Oncogene 8: 2627
2635, 1993
30. Pena SV, Melhem MF, Meisler AI, Cartwright CA:
Elevated c-yes tyrosine kinase activity in premalignant
lesions of the colon. Gastroenterology 108: 117124, 1995
31. Alexander RJ, Panja A, Kaplan-Liss E, Mayer L, Raicht
RF: Expression of protooncogene-encoded mRNA by
colonic epithelial cells in inammatory bowel disease. Dig
Dis Sci 41: 660669, 1996
32. Han NM, Curley SA, Gallick GE: Differential activation
of pp60(c-src) and pp62(c-yes) in human colorectal
carcinoma liver metastases. Clin Cancer Res 2: 1397
1404, 1996
33. Boyce BF, Yoneda T, Lowe C, Soriano P, Mundy GR:
Requirement of pp60c-src expression for osteoclasts to
form rufed borders and resorb bone in mice. J Clin Invest
90: 16221627, 1992

353

34. Ignelzi MA, Jr., Miller DR, Soriano P, Maness PF:


Impaired neurite outgrowth of src-minus cerebellar
neurons on the cell adhesion molecule L1. Neuron 12:
873884, 1994
35. Kaplan KB, Swedlow JR, Morgan DO, Varmus HE: c-Src
enhances the spreading of src/ broblasts on bronectin
by a kinase-independent mechanism. Genes Dev 9: 1505
1517, 1995
36. Hall CL, Lange LA, Prober DA, Zhang S, Turley EA:
pp60(c-src) is required for cell locomotion regulated by the
hyaluronanreceptor RHAMM. Oncogene 13: 22132224,
1996
37. Veillette A, Foss FM, Sausville EA, Bolen JB, Rosen N:
Expression of the lck tyrosine kinase gene in human colon
carcinoma and other non-lymphoid human tumor cell
lines. Oncogene Res 1: 357374, 1987
38. Muise-Helmericks RC, Rosen N: Identication of a novel
repressive element in the proximal lck promoter. J Biol
Chem 270: 2753827543, 1995
39. Irby R, Mao W, Coppola D, Jove R, Gamero A,
Cuthbertson D, Fujita DJ, Yeatman TJ: Overexpression
of normal c-Src in poorly metastatic human colon cancer
cells enhances primary tumor growth but not metastatic
potential. Cell Growth Differ 8: 12871295, 1997
40. Pories SE, Hess DT, Swenson K, Lotz M, Moussa R,
Steele G, Jr., Shibata D, Rieger-Christ KM, Summerhayes
C: Overexpression of pp60c-src elicits invasive behavior in
rat colon epithelial cells. Gastroenterology 114: 1287
1295, 1998
41. Garcia R, Parikh NU, Saya H, Gallick GE: Effect of
herbimycin A on growth and pp60c-src activity in human
colon tumor cell lines. Oncogene 6: 19831989, 1991
42. Kawai N, Tsuji S, Tsujii M, Ito T, Yasumaru M, Kakiuchi
Y, Kimura A, Komori M, Sasaki Y, Hayashi N, Kawano
S, Dubois R, Hori M: Tumor necrosis factor alpha
stimulates invasion of Src-activated intestinal cells.
Gastroenterology 122: 331339, 2002
43. Novotny-Smith CL, Gallick GE: Growth inhibition of
human colorectal carcinoma cell lines by tumor necrosis
factor-alpha correlates with reduced activity of pp60c-src.
J Immunother 11: 159168, 1992
44. Nakagawa T, Tanaka S, Suzuki H, Takayanagi H,
Miyazaki T, Nakamura K, Tsuruo T: Overexpression of
the csk gene suppresses tumor metastasis in vivo. Int J
Cancer 88: 384391, 2000
45. Boyer B, Bourgeois Y, Poupon MF: Src kinase contributes to the metastatic spread of carcinoma cells.
Oncogene 21: 23472356, 2002
46. Staley CA, Parikh NU, Gallick GE: Decreased tumorigenicity of a human colon adenocarcinoma cell line by an
antisense expression vector specic for c-Src. Cell Growth
Differ 8: 269274, 1997
47. Di Domenico M, Castoria G, Bilancio A, Migliaccio A,
Auricchio F: Estradiol activation of human colon
carcinoma-derived Caco-2 cell growth. Cancer Res 56:
45164521, 1996
48. Mao W, Irby R, Coppola D, Fu L, Wloch M, Turner J,
Yu H, Garcia R, Jove R, Yeatman TJ: Activation of c-Src
by receptor tyrosine kinases in human colon cancer cells

354

49.

50.

51.

52.

53.

54.

55.

56.

57.

58.

59.

60.

61.

62.

Summy and Gallick


with high metastatic potential. Oncogene 15: 30833090,
1997
DeSeau V, Rosen N, Bolen JB: Analysis of pp60c-src
tyrosine kinase activity and phosphotyrosyl phosphatase
activity in human colon carcinoma and normal human
colon mucosal cells. J Cell Biochem 35: 113128, 1987
Park J, Cartwright CA: Src activity increases and Yes
activity decreases during mitosis of human colon carcinoma cells. Mol Cell Biol 15: 23742382, 1995
Peng ZY, Cartwright CA: Regulation of the Src tyrosine
kinase and Syp tyrosine phosphatase by their cellular
association. Oncogene 11: 19551962, 1995
Zheng XM, Resnick RJ, Shalloway D: A phosphotyrosine
displacement mechanism for activation of Src by PTPalpha. Embo J 19: 964978, 2000
Cam WR, Masaki T, Shiratori Y, Kato N, Ikenoue T,
Okamoto M, Igarashi K, Sano T, Omata M: Reduced Cterminal Src kinase activity is correlated inversely with
pp60(c-src) activity in colorectal carcinoma. Cancer 92:
6170, 2001
Daigo Y, Furukawa Y, Kawasoe T, Ishiguro H, Fujita M,
Sugai S, Nakamori S, Liefers GJ, Tollenaar RA, van de
Velde CJ, Nakamura Y: Absence of genetic alteration at
codon 531 of the human c-src gene in 479 advanced
colorectal cancers from Japanese and Caucasian patients.
Cancer Res 59: 42224224, 1999
Nilbert M, Fernebro E: Lack of activating c-SRC
mutations at codon 531 in rectal cancer. Cancer Genet
Cytogenet 121: 9495, 2000
Wang NM, Yeh KT, Tsai CH, Chen SJ, Chang JG: No
evidence of correlation between mutation at codon 531 of
src and the risk of colon cancer in Chinese. Cancer Lett
150: 201204, 2000
Rajala RV, Dehm S, Bi X, Bonham K, Sharma RK:
Expression of N-myristoyltransferase inhibitor protein and
its relationship to c-Src levels in human colon cancer cell
lines. Biochem Biophys Res Commun 273: 11161120, 2000
Brunton VG, Ozanne BW, Paraskeva C, Frame MC: A
role for epidermal growth factor receptor, c-Src and focal
adhesion kinase in an in vitro model for the progression of
colon cancer. Oncogene 14: 283293, 1997
Allgayer H, Wang H, Gallick GE, Crabtree A, Mazar A,
Jones T, Kraker AJ, Boyd DD: Transcriptional induction
of the urokinase receptor gene by a constitutively active
Src. Requirement of an upstream motif (-152/-135) bound
with Sp1. J Biol Chem 274: 1842818437, 1999
Owens DW, McLean GW, Wyke AW, Paraskeva C,
Parkinson EK, Frame MC, Brunton VG: The catalytic
activity of the Src family kinases is required to disrupt
cadherin-dependent cell-cell contacts. Mol Biol Cell 11:
5164, 2000
Irby RB, Yeatman TJ: Increased Src activity disrupts
cadherin/catenin-mediated homotypic adhesion in human
colon cancer and transformed rodent cells. Cancer Res 62:
26692674, 2002
Avizienyte E, Wyke AW, Jones RJ, McLean GW,
Westhoff MA, Brunton VG, Frame MC: Src-induced
de-regulation of E-cadherin in colon cancer cells requires
integrin signaling. Nat Cell Biol 4: 632638, 2002

63. Mukhopadhyay D, Tsiokas L, Zhou XM, Foster D,


Brugge JS, Sukhatme VP: Hypoxic induction of human
vascular endothelial growth factor expression through cSrc activation. Nature 375: 577581, 1995
64. Fleming RY, Ellis LM, Parikh NU, Liu W, Staley CA,
Gallick GE: Regulation of vascular endothelial growth
factor expression in human colon carcinoma cells by
activity of src kinase. Surgery 122: 501507, 1997
65. Ellis LM, Staley CA, Liu W, Fleming RY, Parikh NU,
Bucana CD, Gallick GE: Down-regulation of vascular
endothelial growth factor in a human colon carcinoma cell
line transfected with an antisense expression vector
specic for c-src. J Biol Chem 273: 10521057, 1998
66. Malek RL, Irby RB, Guo QM, Lee K, Wong S, He M,
Tsai J, Frank B, Liu ET, Quackenbush J, Jove R,
Yeatman TJ, Lee NH: Identication of Src transformation ngerprint in human colon cancer. Oncogene 21:
72567265, 2002
67. Jacobs C, Rubsamen H: Expression of pp60c-src protein
kinase in adult and fetal human tissue: High activities in
some sarcomas and mammary carcinomas. Cancer Res 43:
16961702, 1983
68. Lehrer S, OShaughnessy J, Song HK, Levine E, Savoretti
P, Dalton J, Lipsztein R, Kalnicki S, Bloomer WD:
Activity of pp60c-src protein kinase in human breast
cancer. Mt Sinai J Med 56: 8385, 1989
69. Koster A, Landgraf S, Leipold A, Sachse R, Gebhart E,
Tulusan AH, Ronay G, Schmidt C, Dingermann T:
Expression of oncogenes in human breast cancer specimens. Anticancer Res 11: 193201, 1991
70. Ottenhoff-Kalff AE, Rijksen G, van Beurden EA,
Hennipman A, Michels AA, Staal GE: Characterization
of protein tyrosine kinases from human breast cancer:
Involvement of the c-src oncogene product. Cancer Res
52: 47734778, 1992
71. Verbeek BS, Vroom TM, Adriaansen-Slot SS, OttenhoffKalff AE, Geertzema JG, Hennipman A, Rijksen G: c-Src
protein expression is increased in human breast cancer. An
immunohistochemical and biochemical analysis. J Pathol
180: 383388, 1996
72. Reissig D, Clement J, Sanger J, Berndt A, Kosmehl H,
Bohmer FD: Elevated activity and expression of Srcfamily kinases in human breast carcinoma tissue versus
matched non-tumor tissue. J Cancer Res Clin Oncol 127:
226230, 2001
73. Biscardi JS, Ishizawar RC, Silva CM, Parsons SJ:
Tyrosine kinase signaling in breast cancer: Epidermal
growth factor receptor and c-Src interactions in breast
cancer. Breast Cancer Res 2: 203210, 2000
74. Alonso G, Koegl M, Mazurenko N, Courtneidge SA:
Sequence requirements for binding of Src family tyrosine
kinases to activated growth factor receptors. J Biol Chem
270: 98409848, 1995
75. Courtneidge SA, Dhand R, Pilat D, Twamley GM,
Watereld MD, Roussel MF: Activation of Src family
kinases by colony stimulating factor-1, and their association with its receptor. Embo J 12: 943950, 1993
76. Kypta RM, Goldberg Y, Ulug ET, Courtneidge SA:
Association between the PDGF receptor and members

Src family kinases in tumor progression and metastasis

77.

78.

79.

80.

81.

82.

83.

84.

85.

86.

87.

88.

89.

of the src family of tyrosine kinases. Cell 62: 481492,


1990
Mori S, Ronnstrand L, Yokote K, Engstrom A,
Courtneidge SA, Claesson-Welsh L, Heldin CH: Identication of two juxtamembrane autophosphorylation sites
in the PDGF beta-receptor; involvement in the interaction
with Src family tyrosine kinases. Embo J 12: 22572264,
1993
Twamley GM, Kypta RM, Hall B, Courtneidge SA:
Association of Fyn with the activated platelet-derived
growth factor receptor: Requirements for binding and
phosphorylation. Oncogene 7: 18931901, 1992
Luttrell DK, Lee A, Lansing TJ, Crosby RM, Jung KD,
Willard D, Luther M, Rodriguez M, Berman J, Gilmer
TM: Involvement of pp60c-src with two major signaling
pathways in human breast cancer. Proc Natl Acad Sci
USA 91: 8387, 1994
Oude Weernink PA, Ottenhoff-Kalff AE, Vendrig MP,
van Beurden EA, Staal GE, Rijksen G: Functional
interaction between the epidermal growth factor receptor
and c-Src kinase activity. FEBS Lett 352: 296300, 1994
Muthuswamy SK, Muller WJ: Activation of the Src
family of tyrosine kinases in mammary tumorigenesis.
Adv Cancer Res 64: 111123, 1994
Muthuswamy SK, Siegel PM, Dankort DL, Webster MA,
Muller WJ: Mammary tumors expressing the neu protooncogene possess elevated c-Src tyrosine kinase activity.
Mol Cell Biol 14: 735743, 1994
Muthuswamy SK, Muller WJ: Direct and specic interaction of c-Src with Neu is involved in signaling by the
epidermal growth factor receptor. Oncogene 11: 271279,
1995
Muthuswamy SK, Muller WJ: Activation of Src family
kinases in Neu-induced mammary tumors correlates with
their association with distinct sets of tyrosine phosphorylated proteins in vivo. Oncogene 11: 18011810, 1995
Biscardi JS, Belsches AP, Parsons SJ: Characterization of
human epidermal growth factor receptor and c-Src
interactions in human breast tumor cells. Mol Carcinog
21: 261272, 1998
Li Y, Ren J, Yu W, Li Q, Kuwahara H, Yin L, Carraway
KL, 3rd, Kufe D: The epidermal growth factor receptor
regulates interaction of the human DF3/MUC1 carcinoma antigen with c-Src and beta-catenin. J Biol Chem
276: 3523935242, 2001
Egan C, Pang A, Durda D, Cheng HC, Wang JH, Fujita
DJ: Activation of Src in human breast tumor cell lines:
Elevated levels of phosphotyrosine phosphatase activity
that preferentially recognizes the Src carboxy terminal
negative regulatory tyrosine 530. Oncogene 18: 1227
1237, 1999
Bjorge JD, Pang A, Fujita DJ: Identication of proteintyrosine phosphatase 1B as the major tyrosine phosphatase activity capable of dephosphorylating and activating
c-Src in several human breast cancer cell lines. J Biol
Chem 275: 4143941446, 2000
Webster MA, Cardiff RD, Muller WJ: Induction of
mammary epithelial hyperplasias and mammary tumors in
transgenic mice expressing a murine mammary tumor

90.

91.

92.

93.

94.

95.

96.

97.

98.

99.

100.

101.

102.

355

virus/activated c-src fusion gene. Proc Natl Acad Sci USA


92: 78497853, 1995
Guy CT, Muthuswamy SK, Cardiff RD, Soriano P,
Muller WJ: Activation of the c-Src tyrosine kinase is
required for the induction of mammary tumors in
transgenic mice. Genes Dev 8: 2332, 1994
Arnold SF, Obourn JD, Jaffe H, Notides AC: Phosphorylation of the human estrogen receptor on tyrosine 537 in
vivo and by src family tyrosine kinases in vitro. Mol
Endocrinol 9: 2433, 1995
Arnold SF, Vorojeikina DP, Notides AC: Phosphorylation of tyrosine 537 on the human estrogen receptor is
required for binding to an estrogen response element.
J Biol Chem 270: 3020530212, 1995
Migliaccio A, Di Domenico M, Castoria G, de Falco A,
Bontempo P, Nola E, Auricchio F: Tyrosine kinase/
p21ras/MAP-kinase pathway activation by estradiolreceptor complex in MCF-7 cells. Embo J 15: 1292
1300, 1996
Migliaccio A, Piccolo D, Castoria G, Di Domenico M,
Bilancio A, Lombardi M, Gong W, Beato M, Auricchio
F: Activation of the Src/p21ras/Erk pathway by progesterone receptor via cross-talk with estrogen receptor.
Embo J 17: 20082018, 1998
Boonyaratanakornkit V, Scott MP, Ribon V, Sherman L,
Anderson SM, Maller JL, Miller WT, Edwards DP:
Progesterone receptor contains a proline-rich motif that
directly interacts with SH3 domains and activates c-Src
family tyrosine kinases. Mol Cell 8: 269280, 2001
Castoria G, Migliaccio A, Bilancio A, Di Domenico M, de
Falco A, Lombardi M, Fiorentino R, Varricchio L,
Barone MV, Auricchio F: PI3-kinase in concert with Src
promotes the S-phase entry of oestradiol-stimulated
MCF-7 cells. Embo J 20: 60506059, 2001
Tsai EM, Wang SC, Lee JN, Hung MC: Akt activation by
estrogen in estrogen receptor-negative breast cancer cells.
Cancer Res 61: 83908392, 2001
Maa MC, Leu TH, McCarley DJ, Schatzman RC,
Parsons SJ: Potentiation of epidermal growth factor
receptor-mediated oncogenesis by c-Src: Implications for
the etiology of multiple human cancers. Proc Natl Acad
Sci USA 92: 69816985, 1995
Shefeld LG: C-Src activation by ErbB2 leads to
attachment-independent growth of human breast epithelial cells. Biochem Biophys Res Commun 250: 2731, 1998
Amundadottir LT, Leder P: Signal transduction pathways
activated and required for mammary carcinogenesis in
response to specic oncogenes. Oncogene 16: 737746,
1998
Vercoutter-Edouart A, Lemoine J, Smart CE, Nurcombe
V, Boilly B, Peyrat J, Hondermarck H: The mitogenic
signaling pathway for broblast growth factor-2 involves
the tyrosine phosphorylation of cyclin D2 in MCF-7
human breast cancer cells. FEBS Lett 478: 209215,
2000
Belsches-Jablonski AP, Biscardi JS, Peavy DR, Tice DA,
Romney DA, Parsons SJ: Src family kinases and HER2
interactions in human breast cancer cell growth and
survival. Oncogene 20: 14651475, 2001

356

Summy and Gallick

103. Moasser MM, Srethapakdi M, Sachar KS, Kraker AJ,


Rosen N: Inhibition of Src kinases by a selective tyrosine
kinase inhibitor causes mitotic arrest. Cancer Res 59:
61456152, 1999
104. Hung W, Elliott B: Co-operative effect of c-Src tyrosine
kinase and Stat3 in activation of hepatocyte growth factor
expression in mammary carcinoma cells. J Biol Chem 276:
1239512403, 2001
105. Maulik G, Shrikhande A, Kijima T, Ma PC, Morrison
PT, Salgia R: Role of the hepatocyte growth factor
receptor, c-Met, in oncogenesis and potential for therapeutic inhibition. Cytokine Growth Factor Rev 13: 4159,
2002
106. Garcia R, Bowman TL, Niu G, Yu H, Minton S, MuroCacho CA, Cox CE, Falcone R, Fairclough R, Parsons S,
Laudano A, Gazit A, Levitzki A, Kraker A, Jove R:
Constitutive activation of Stat3 by the Src and JAK
tyrosine kinases participates in growth regulation of
human breast carcinoma cells. Oncogene 20: 24992513,
2001
107. Olayioye MA, Badache A, Daly JM, Hynes NE: An
essential role for Src kinase in ErbB receptor signaling
through the MAPK pathway. Exp Cell Res 267: 8187,
2001
108. Bougeret C, Jiang S, Keydar I, Avraham H: Functional
analysis of Csk and CHK kinases in breast cancer cells.
J Biol Chem 276: 3371133720, 2001
109. McShan GD, Zagozdzon R, Park SY, Zrihan-Licht S, Fu
Y, Avraham S, Avraham H: Csk homologous kinase
associates with RAFTK/Pyk2 in breast cancer cells and
negatively regulates its activation and breast cancer cell
migration. Int J Oncol 21: 197205, 2002
110. Pal S, Datta K, Mukhopadhyay D: Central role of p53 on
regulation of vascular permeability factor/vascular
endothelial growth factor (VPF/VEGF) expression in
mammary carcinoma. Cancer Res 61: 69526957,
2001
111. Hall CL, Turley EA: Hyaluronan: RHAMM mediated cell
locomotion and signaling in tumorigenesis. J Neurooncol
26: 221229, 1995
112. Zrihan-Licht S, Fu Y, Settleman J, Schinkmann K, Shaw
L, Keydar I, Avraham S, Avraham H: RAFTK/Pyk2
tyrosine kinase mediates the association of p190 RhoGAP
with RasGAP and is involved in breast cancer cell
invasion. Oncogene 19: 13181328, 2000
113. Gutwein P, Oleszewski M, Mechtersheimer S, AgmonLevin N, Krauss K, Altevogt P: Role of Src kinases in
the ADAM-mediated release of L1 adhesion molecule
from human tumor cells. J Biol Chem 275: 1549015497,
2000
114. Rahimi N, Hung W, Tremblay E, Saulnier R, Elliott B: cSrc kinase activity is required for hepatocyte growth
factor-induced motility and anchorage-independent
growth of mammary carcinoma cells. J Biol Chem 273:
3371433721, 1998
115. Nam JS, Ino Y, Sakamoto M, Hirohashi S: Src family
kinase inhibitor PP2 restores the E-Cadherin/Catenin cell
adhesion system in human cancer cells and reduces cancer
metastasis. Clin Cancer Res 8: 24302436, 2002

116. Lee HJ, Kim E, Jee B, Hahn JH, Han K, Jung KC, Park
SH, Lee H: Functional involvement of src and focal
adhesion kinase in a CD99 splice variant-induced motility
of human breast cancer cells. Exp Mol Med 34: 177183,
2002
117. Loganzo F Jr., Dosik JS, Zhao Y, Vidal MJ, Nanus DM,
Sudol M, Albino AP: Elevated expression of protein
tyrosine kinase c-Yes, but not c-Src, in human malignant
melanoma. Oncogene 8: 26372644, 1993
118. Marchetti D, Parikh N, Sudol M, Gallick GE: Stimulation
of the protein tyrosine kinase c-Yes but not c-Src by
neurotrophins in human brain-metastatic melanoma cells.
Oncogene 16: 32533260, 1998
119. Visser CJ, Rijksen G, Woutersen RA, De Weger RA:
Increased immunoreactivity and protein tyrosine kinase
activity of the protooncogene pp60c-src in preneoplastic
lesions in rat pancreas. Lab Invest 74: 211, 1996
120. Lutz MP, Esser IB, Flossmann-Kast BB, Vogelmann R,
Luhrs H, Friess H, Buchler MW, Adler G: Overexpression
and activation of the tyrosine kinase Src in human
pancreatic carcinoma. Biochem Biophys Res Commun
243: 503508, 1998
121. MacMillan-Crow LA, Greendorfer JS, Vickers SM,
Thompson JA: Tyrosine nitration of c-SRC tyrosine
kinase in human pancreatic ductal adenocarcinoma.
Arch Biochem Biophys 377: 350356, 2000
122. Flossmann-Kast BB, Jehle PM, Hoeich A, Adler G, Lutz
MP: Src stimulates insulin-like growth factor I (IGF-I)dependent cell proliferation by increasing IGF-I receptor
number in human pancreatic carcinoma cells. Cancer Res
58: 35513554, 1998
123. Tanno S, Mitsuuchi Y, Altomare DA, Xiao GH, Testa
JR: AKT activation up-regulates insulin-like growth
factor I receptor expression and promotes invasiveness
of human pancreatic cancer cells. Cancer Res 61: 589593,
2001
124. Menke A, Philippi C, Vogelmann R, Seidel B, Lutz MP,
Adler G, Wedlich D: Down-regulation of E-cadherin gene
expression by collagen type I and type III in pancreatic
cancer cell lines. Cancer Res 61: 35083517, 2001
125. Wiener JR, Nakano K, Kruzelock RP, Bucana CD, Bast
RC Jr., Gallick GE: Decreased Src tyrosine kinase activity
inhibits malignant human ovarian cancer tumor growth in
a nude mouse model. Clin Cancer Res 5: 21642170, 1999
126. Venkatakrishnan G, Salgia R, Groopman JE: Chemokine
receptors CXCR-1/2 activate mitogen-activated protein
kinase via the epidermal growth factor receptor in ovarian
cancer cells. J Biol Chem 275: 68686875, 2000
127. Bourguignon LY, Zhu H, Shao L, Chen YW: CD44
interaction with c-Src kinase promotes cortactin-mediated
cytoskeleton function and hyaluronic acid-dependent
ovarian tumor cell migration. J Biol Chem 276: 7327
7336, 2001
128. Fanning P, Bulovas K, Saini KS, Libertino JA, Joyce AD,
Summerhayes IC: Elevated expression of pp60c-src in low
grade human bladder carcinoma. Cancer Res 52: 1457
1462, 1992
129. Rodier JM, Valles AM, Denoyelle M, Thiery JP, Boyer B:
pp60c-src is a positive regulator of growth factor-induced

Src family kinases in tumor progression and metastasis

130.

131.

132.

133.

134.

135.

136.

137.

138.

139.

140.

141.

142.

cell scattering in a rat bladder carcinoma cell line. J Cell


Biol 131: 761773, 1995
Cattan N, Rochet N, Mazeau C, Zanghellini E, Mari B,
Chauzy C, Stora de Novion H, Amiel J, Lagrange JL,
Rossi B, Gioanni J: Establishment of two new human
bladder carcinoma cell lines, CAL 29 and CAL 185.
Comparative study of cell scattering and epithelial to
mesenchyme transition induced by growth factors. Br J
Cancer 85: 14121417, 2001
Takekura N, Yasui W, Yoshida K, Tsujino T, Nakayama
H, Kameda T, Yokozaki H, Nishimura Y, Ito H, Tahara
E: pp60c-src protein kinase activity in human gastric
carcinomas. Int J Cancer 45: 847851, 1990
Jankowski J, Coghill G, Hopwood D, Wormsley KG:
Oncogenes and onco-suppressor gene in adenocarcinoma
of the oesophagus. Gut 33: 10331038, 1992
Kumble S, Omary MB, Cartwright CA, Triadalopoulos
G: Src activation in malignant and premalignant epithelia
of Barretts esophagus. Gastroenterology 112: 348356,
1997
van Oijen MG, Rijksen G, ten Broek FW, Slootweg PJ:
Overexpression of c-Src in areas of hyperproliferation in
head and neck cancer, premalignant lesions and benign
mucosal disorders. J Oral Pathol Med 27: 147152,
1998
Kiefer PE, Wegmann B, Bacher M, Erbil C, Heidtmann
H, Havemann K: Different pattern of expression of
cellular oncogenes in human non-small-cell lung cancer
cell lines. J Cancer Res Clin Oncol 116: 2937, 1990
Mazurenko NN, Kogan EA, Zborovskaya IB, Kisseljov
FL: Expression of pp60c-src in human small cell and nonsmall cell lung carcinomas. Eur J Cancer 28: 372377,
1992
Budde RJ, Ke S, Levin VA: Activity of pp60c-src in 60
different cell lines derived from human tumors. Cancer
Biochem Biophys 14: 171175, 1994
Krystal GW, DeBerry CS, Linnekin D, Litz J: Lck
associates with and is activated by Kit in a small cell lung
cancer cell line: Inhibition of SCF-mediated growth by the
Src family kinase inhibitor PP1. Cancer Res 58: 4660
4666, 1998
Bondzi C, Litz J, Dent P, Krystal GW: Src family kinase
activity is required for Kit-mediated mitogen-activated
protein (MAP) kinase activation, however loss of functional retinoblastoma protein makes MAP kinase activation unnecessary for growth of small cell lung cancer cells.
Cell Growth Differ 11: 305314, 2000
Sato M, Tanaka T, Maeno T, Sando Y, Suga T, Maeno
Y, Sato H, Nagai R, Kurabayashi M: Inducible expression of endothelial PAS domain protein-1 by hypoxia in
human lung adenocarcinoma A549 cells. Role of Src
family kinases-dependent pathway. Am J Respir Cell Mol
Biol 26: 127134, 2002
Takenaka N, Mikoshiba K, Takamatsu K, Tsukada Y,
Ohtani M, Toya S: Immunohistochemical detection of the
gene product of Rous sarcoma virus in human brain
tumors. Brain Res 337: 201207, 1985
Bolen JB, Rosen N, Israel MA: Increased pp60c-src
tyrosyl kinase activity in human neuroblastomas is

143.

144.

145.

146.

147.

148.

149.

150.

151.

152.

153.

154.

155.

357

associated with amino-terminal tyrosine phosphorylation


of the src gene product. Proc Natl Acad Sci USA 82:
72757279, 1985
OShaughnessy J, Deseau V, Amini S, Rosen N, Bolen JB:
Analysis of the c-src gene product structure, abundance,
and protein kinase activity in human neuroblastoma and
glioblastoma cells. Oncogene Res 2: 118, 1987
Mellstrom K, Bjelfman C, Hammerling U, Pahlman S:
Expression of c-src in cultured human neuroblastoma and
small-cell lung carcinoma cell lines correlates with
neurocrine differentiation. Mol Cell Biol 7: 41784184,
1987
Pahlman S, Hammerling U: Src expression in small-cell
lung carcinoma and other neuroendocrine malignancies.
Am Rev Respir Dis 142: S5456, 1990
Bjelfman C, Hedborg F, Johansson I, Nordenskjold M,
Pahlman S: Expression of the neuronal form of pp60c-src
in neuroblastoma in relation to clinical stage and
prognosis. Cancer Res 50: 69086914, 1990
Weissenberger J, Steinbach JP, Malin G, Spada S, Rulicke
T, Aguzzi A: Development and malignant progression of
astrocytomas in GFAP-v-src transgenic mice. Oncogene
14: 20052013, 1997
Theurillat JP, Hainfellner J, Maddalena A, Weissenberger
J, Aguzzi A: Early induction of angiogenetic signals in
gliomas of GFAP-v-src transgenic mice. Am J Pathol 154:
581590, 1999
Hecker TP, Grammer JR, Gillespie GY, Stewart J Jr.,
Gladson CL: Focal adhesion kinase enhances signaling
through the Shc/extracellular signal-regulated kinase
pathway in anaplastic astrocytoma tumor biopsy samples.
Cancer Res 62: 26992707, 2002
Torigoe T, OConnor R, Santoli D, Reed JC: Interleukin3 regulates the activity of the LYN protein-tyrosine kinase
in myeloid-committed leukemic cell lines. Blood 80: 617
624, 1992
Hallek M, Neumann C, Schaffer M, Danhauser-Riedl S,
von Bubnoff N, de Vos G, Druker BJ, Yasukawa K,
Grifn JD, Emmerich B: Signal transduction of interleukin-6 involves tyrosine phosphorylation of multiple
cytosolic proteins and activation of Src-family kinases
Fyn, Hck, and Lyn in multiple myeloma cell lines. Exp
Hematol 25: 13671377, 1997
Ishikawa H, Tsuyama N, Abroun S, Liu S, Li FJ,
Taniguchi O, Kawano MM: Requirements of src family
kinase activity associated with CD45 for myeloma cell
proliferation by interleukin-6. Blood 99: 21722178, 2002
Choi SH, Yamanashi Y, Shiota M, Takanashi M, Hojo I,
Itoh T, Watanabe T, Yamamoto T, Mori S: Expression of
Lyn protein on human malignant lymphomas. Lab Invest
69: 736742, 1993
Myers DE, Jun X, Waddick KG, Forsyth C, Chelstrom
LM, Gunther RL, Tumer NE, Bolen J, Uckun FM:
Membrane-associated CD19-LYN complex is an endogenous p53-independent and Bc12-independent regulator
of apoptosis in human B-lineage lymphoma cells. Proc
Natl Acad Sci USA 92: 95759579, 1995
Holland J, Owens T: Signaling through intercellular
adhesion molecule 1 (ICAM-1) in a B cell lymphoma

358

156.

157.

158.

159.

160.

161.

162.

163.

Summy and Gallick


line. The activation of Lyn tyrosine kinase and the
mitogen-activated protein kinase pathway. J Biol Chem
272: 91089112, 1997
Katagiri K, Yokoyama KK, Yamamoto T, Omura S, Irie
S, Katagiri T: Lyn and Fgr protein-tyrosine kinases
prevent apoptosis during retinoic acid-induced granulocytic differentiation of HL-60 cells. J Biol Chem 271:
1155711562, 1996
Roginskaya V, Zuo S, Caudell E, Nambudiri G, Kraker AJ,
Corey SJ: Therapeutic targeting of Src-kinase Lyn in
myeloid leukemic cell growth. Leukemia 13: 855861, 1999
Tilbrook PA, Palmer GA, Bittorf T, McCarthy DJ,
Wright MJ, Sarna MK, Linnekin D, Cull VS, Williams
JH, Ingley E, Schneider-Mergener J, Krystal G, Klinken
SP: Maturation of erythroid cells and erythroleukemia
development are affected by the kinase activity of Lyn.
Cancer Res 61: 24532458, 2001
Ptasznik A, Urbanowska E, Chinta S, Costa MA, Katz
BA, Stanislaus MA, Demir G, Linnekin D, Pan ZK,
Gewirtz AM: Crosstalk between BCR/ABL oncoprotein
and CXCR4 signaling through a Src family kinase in
human leukemia cells. J Exp Med 196: 667678, 2002
Lionberger JM, Wilson MB, Smithgall TE: Transformation of myeloid leukemia cells to cytokine independence
by Bcr-Abl is suppressed by kinase-defective Hck. J Biol
Chem 275: 1858118585, 2000
Penninger JM, Wallace VA, Kishihara K, Mak TW: The
role of p56lck and p59fyn tyrosine kinases and CD45
protein tyrosine phosphatase in T-cell development and
clonal selection. Immunol Rev 135: 183214, 1993
Perlmutter RM: Control of T cell development by nonreceptor protein tyrosine kinases. Cancer Surv 22: 8595,
1995
Sefton BM: The lck tyrosine protein kinase. Oncogene 6:
683686, 1991

164. Sefton BM, Taddie JA: Role of tyrosine kinases in


lymphocyte activation. Curr Opin Immunol 6: 372379,
1994
165. Abts H, Jucker M, Diehl V, Tesch H: Human chronic
lymphocytic leukemia cells regularly express mRNAs of
the protooncogenes lck and c-fgr. Leuk Res 15: 987997,
1991
166. Waddick KG, Chae HP, Tuel-Ahlgren L, Jarvis LJ,
Dibirdik I, Myers DE, Uckun FM: Engagement of
the CD19 receptor on human B-lineage leukemia
cells activates LCK tyrosine kinase and facilitates
radiation-induced apoptosis. Radiat Res 136: 313319,
1993
167. Miyazaki T, Liu ZJ, Taniguchi T: Selective cooperation of
HTLV-1-encoded p40tax-1 with cellular oncoproteins in
the induction of hematopoietic cell proliferation. Oncogene 12: 24032408, 1996
168. Majolini MB, DElios MM, Galieni P, Boncristiano M,
Lauria F, Del Prete G, Telford JL, Baldari CT: Expression
of the T-cell-specic tyrosine kinase Lck in normal B-1
cells and in chronic lymphocytic leukemia B cells. Blood
91: 33903396, 1998
169. Metcalf CA III, van Schravendijk MR, Dalgarno DC,
Sawyer TK: Targeting protein kinases for bone disease:
Discovery and development of SRC inhibitors. Curr
Pharm Des 8: 20492075, 2002
170. Windham TC, Parikh NU, Siwak DR, Summy JM,
McConkey DJ, Kraker AJ, Gallick GE: Src activation
regulates anoikis in human colon tumor cell lines.
Oncogene 21: 77977807, 2002
171. Donato NJ, Wu JY, Stapley J, Gallick G, Lin H,
Arlinghaus R, Talpaz M: BCR-ABL independence and
LYN kinase overexpression in chronic myelogenous
leukemia cells selected for resistance to STI571. Blood
101: 690698, 2003

Vous aimerez peut-être aussi