Вы находитесь на странице: 1из 16

Endocrine-Related Cancer (1999) 6 389-404

Prolactin involvement in breast


cancer
B K Vonderhaar
Molecular and Cellular Endocrinology Section, Laboratory of Tumor Immunology and Biology, National Cancer
Institute, National Institutes of Health, Bethesda, Maryland, USA
(Requests for offprints should be addressed to B K Vonderhaar, NCI, NIH, 10 Center Drive, Building 10, Room
5B47, Bethesda, Maryland 20892-1402, USA)

Abstract
Normal development and differentiation of the mammary gland are profoundly influenced by prolactin
(PRL). In rodent mammary cancer PRL plays a well defined role, but its role in human breast cancer
has not been appreciated until recently. It is now clear that breast tissue, both normal and malignant,
is a significant source of extrapituitary PRL. Thus an autocrine/paracrine role of PRL in human breast
cancer may be invoked. Both PRL and PRL receptor mRNA are expressed in the vast majority of
breast cancer biopsies independent of estrogen and progesterone receptor status. An autocrine/
paracrine PRL acting in human breast cancer requires that this hormones action be blocked at the
cellular level, as opposed to suppressing the synthesis and secretion of pituitary PRL. Mutants of PRL
or human growth hormone are being explored which act as selective PRL antagonists. In addition,
tamoxifen has been shown to act locally at the target tissue by binding directly to the PRL receptor
and thus inhibiting PRLs action. These strategies may have clinical relevance in treating PRLresponsive human breast cancer.
Endocrine-Related Cancer (1999) 6 389-404

Breast cancer models and prolactin


Historically, the role of the peptide hormone prolactin
(PRL) in human breast cancer has been an obscure one. It
has been, and still is, difficult to assign a role of this
hormone in either the etiology or progression of the disease. However, it is well established that PRL is intimately
involved in development and differentiation of the normal
gland in mammalian species, and in rodent model systems
it is a key player in the development of mammary cancer.
Much of the literature on human breast cancer and PRL
appears to be contradictory. Thus, it has been difficult to
establish a defined involvement of PRL in human breast
disease. However, it is hard to believe that a hormone so
intimately involved in normal breast growth and
differentiation would have nothing to do with breast
malignancy. This paper examines the basis for a claim that
PRL is a player in human breast cancer and that it is
through an autocrine/paracrine pathway that this hormone
acts.
Rodent mammary cancer
In the normal mammary gland, growth, development and
differentiation is influenced by a variety of growth factors
and hormones; the peptide hormone, prolactin, is consid-

ered a major player (Vonderhaar 1987). Functional differentiation of the gland as measured by induction of milk
protein synthesis both in vivo and in vitro is dependent on
PRL. Extensive growth and development of the alveolar
cells within the lobules require PRL (Vonderhaar 1984,
1987). Progesterone increases the level of PRL receptors
in the developing mammary gland (Nagasawa et al. 1985)
and thus acts synergistically with PRL to induce cell
growth.
Prolactins role in rodent mammary cancer has been
established for some time (Welsch & Nagasawa 1977,
Vonderhaar & Biswas 1987). Multiple pituitary isografts
which secrete large amounts of PRL result in a significant
increase in the incidence of spontaneous mammary tumors
in mice (Muhlbock & Boot 1959). Daily injections of PRL
result in more mice with tumors compared with controls
(Boot et al. 1962). In rats, there is a direct correlation
between serum PRL levels and susceptibility of various rat
strains to induction of mammary tumors by chemical
carcinogens (Boyns et al. 1973). Tumors induced in rats
by either nitrosomethyl urea or 7,12-dimethyl-benz[a]anthracene are dependent on PRL for sustained growth
(Mershon et al. 1995). In rodents there is a direct
correlation between drug-induced hyperprolactinemia and

Endocrine-Related Cancer (1999) 6 389-404


1351-0088/99/006-389 1999 Society for Endocrinology Printed in Great Britain

Online version via http://www.endocrinology.org

Vonderhaar: Prolactin and breast cancer


increased tumor growth, and hypoprolactinemia and retarded tumor growth (Welsch & Gribler 1973, Welsch &
Nagasawa 1977).
Human breast cancer
In contrast to the situation in rodent model systems, the
function of PRL in the etiology and progression of human
breast cancer is not clear. The discovery of extrapituitary
PRL synthesis by breast tissue suggests that a re-examination of the role of PRL in human breast cancer is in
order. For several years significant contradictory evidence in the literature on the role of this hormone in human
disease has clouded the situation. However, there is
significant evidence that PRL may play a role in human
breast disease. In 1984, Holtkamp et al. (1984) reported
that as many as 44% of patients with metastatic breast
disease were hyperprolactinemic during the course of the
disease. Several cases of breast carcinoma in association
with prolactinoma have been reported (Strung et al. 1997).
In a subset of women at risk for familial breast cancer,
basal serum PRL levels were significantly elevated (Love
& Rose 1985, Love et al. 1991). In addition, the circadian
rhythm of PRL secretion from the pituitary differs
between groups at high vs low risk of breast cancer (Haus
et al. 1980), with no seasonal variations (Holdaway et al.
1997). In one study, hyperprolactinemia was found to be
an important indicator of unfavorable prognosis in nodepositive breast cancer patients, both when evaluated singly
and in conjunction with steroid receptor status
(Bhatavdekar et al. 1994). In another study (Patel et al.
1996), hyperprolactinemia and alterations in levels of p53
were associated with aggressiveness of the tumor, early
disease relapse or metastases, and poor overall survival in
patients with node-negative breast cancer. However, in
contrast, a surgery-induced rise in PRL was paradoxically
associated with a longer disease-free survival in operable
breast carcinoma in patients both with or without axillary
node involvement, despite the potential stimulation of
cancer cell growth by the hormone (Lissoni et al. 1995).
In addition, a significant decline in the serum level of
insulin-like growth factor (IGF)-I was associated with
surgery-induced hyperprolactinemia (Barni et al. 1994),
suggesting that the balance of specific hormones and
growth factors may be a key etiological factor.
Prolactin effects in vivo
It has been known for nearly 20 years that more than 70%
of human breast biopsies are positive for PRL receptors
(Codegone et al. 1981, Peyrat et al. 1981, Bonneterre et al.
1982, LHermite-Baleriaux et al. 1984); approximately
80% of breast cancer cells in culture respond to PRLs
mitogenic signal when proper conditions of reduced
serum or serum-free conditions are employed (Das &
Vonderhaar 1997a). PRL itself can be detected in 60-85%

390

of human breast cancer biopsies by immuno-histochemistry (Purnell et al. 1982, Agarwal et al. 1989).
However, while the literature suggests that PRL levels
may influence human breast cancer, there is no clear correlation between circulating PRL levels and the etiology or
prognosis of the disease (Wang et al. 1986, Ingram et al.
1990, Love et al. 1991). Nor is there a clear clinical
response when the secretion of pituitary PRL is impaired.
In earlier studies, treating patients with PRL inhibiting
ergot drugs in order to diminish the levels of circulating
pituitary PRL resulted in no change in disease state over
time (Henson et al. 1972, Pearson & Manni 1978).
However, human growth hormone (hGH), which is also a
lactogen, may have compensated for the diminished PRL
levels. Therefore, another cohort of women with advanced
breast cancer was given a combination therapy of
bromocriptine, to diminish circulating PRL levels, and a
somatostatin analog to block hGH action (Manni et al.
1989). As a result, circulating levels of PRL, detected by
a single assay, were abolished nearly completely in 8 of 9
patients, whereas hGH levels were suppressed in 7 of 9
patients during treatment. However, the patients entering
the study had been pretreated heavily with chemotherapeutic agents, thus making reliable assessment of
overall antitumor effects difficult. Only one patient
experienced disease stabilization. In a subsequent study, 4
of 6 patients with advanced breast cancer who had failed
first and second line endocrine therapies experienced no
evidence of disease progression for periods of up to 6
months when treated for extended periods of time with
bromocriptine and the long-acting, superpotent somatostatin analog octreotide. While 24 h profiles of immunoreactive PRL, growth hormone (GH) and IGF-I in their
serum were greatly reduced by these treatments, GH levels
and diurnal peaks of bioreactive PRL were still apparent,
although much reduced (Anderson et al. 1993).
Prolactin effects in vitro
Investigations in vitro using human breast tumor tissue
and cells show clear responses to PRL in contrast to results
collected in vivo. Among the responses are increased
synthesis of DNA (Salih et al. 1972, Welsch et al. 1976,
Peyrat et al. 1984, Calaf et al. 1986), protein (Burke &
Gaffney 1978) and -lactalbumin (Wilson et al. 1980),
colony formation (Malarkey et al. 1983, Manni et al.
1986) and changes in shape, adhesion, lipid accumulation
(Shiu & Paterson 1984) and estrogen receptor (ER)
content (Shafie & Grantham 1981). Increased growth is
observed when primary breast biopsy samples are grown
in nude mice supplemented with PRL (McManus &
Welsch 1984). Physiological levels of human PRL (hPRL)
and hGH in culture increase the population doubling of
primary breast tumor biopsies (Malarkey et al. 1983). In
addition, we have shown that more than 80% of ER-

Endocrine-Related Cancer (1999) 6 389-404

Table 1 Synthesis of PRL as a function of ER and PRL receptor (PRL-R) status.


Reprinted from Vonderhaar (1999).

Cell line

ER status

PRL-R status

PRL mRNA

ZR 75-1

T47D

MCF-7

T47Dco

MCF 10a

BT20

MDA MB-231

MDA MB-468

MDA MB-453

Hs578t

SK-BR-3

Not done

A1N4

positive, as well as ER-negative, human breast cancer cell


lines express mRNA for PRL receptors, bind PRL
specifically with high affinity, and respond to PRLs
mitogenic signal (Biswas & Vonderhaar 1987, Das &
Vonderhaar 1997a). Both T47D and MCF-7 human breast
cancer cells respond to PRLs growth signal when grown
as solid tumors in nude mice (Vonderhaar & Biswas 1987).
In the past, several reports have shown a lack of growth
stimulation by PRL with these same human breast cancer
cells. However, these studies were performed in the
presence of high levels of fetal bovine serum, a rich source
of lactogenic hormones (Biswas & Vonderhaar 1987).
Only under serum-free conditions or in the presence of
charcoal-stripped serum (CSS) can direct effects of PRL
on growth of human breast cancer cells be achieved (Das
et al. 1994, Lemus-Wilson et al. 1995, Das & Vonderhaar
1997a). We showed that PRL-stimulated growth of MCF7 cells is greater in the presence of 1% CSS compared with
10% CSS or fetal bovine serum (Biswas & Vonderhaar
1987). Growth effects are seen at concentrations as low as
25 ng/ml hPRL; the maximal effect is observed at 100-250
ng/ml. Other lactogens such as hGH, human placental
lactogen and ovine PRL also stimulate growth, but higher
concentrations are required to achieve the same effect.
Bovine PRL had no effect on the growth of these human
breast cancer cells. Thus MCF-7 cells are more sensitive
to the mitogenic effect of hPRL than to other lactogens
or PRL from other species.
The effects of PRL on the growth of breast cancer cells
is affected by the presence of other hormones and growth
factors in the media. Growth of MCF-7 and ZR75.1 cells
induced by hPRL is completely blocked by melatonin, the
primary hormone from the pineal gland (Lemus-Wilson et

al. 1995). When bovine PRL is added simultaneously with


hPRL to the culture media the effect of hPRL on the
growth of MCF-7 cells is completely abolished. As little
as 50-100 ng/ml bovine PRL is able to block the hPRLinduced growth of these cells (Biswas & Vonderhaar 1987,
Das & Vonderhaar 1997a). In contrast, bovine PRL is an
effective mitogen and differentiating agent in normal
mouse mammary cell lines in culture, suggesting that this
hormones action as an antagonist of hPRL may be unique
to human breast cancer cells.

Synthesis and secretion of prolactin


by breast cancer cells
In human breast cancer, the lack of clear clinical data to
correlate with responses to PRL observed in vitro may be
explained, in part, by the fact that breast cancer cells
themselves are a source of extrapituitary PRL (Fig. 1). For
many years, evidence in the literature has hinted at the
possibility of an extrapituitary source of PRL. In one
study, the levels of PRL in serum remained at 30-80% of
the pre-surgical levels for as long as 10 months in breast
cancer patients following total hypophysectomy (Lachelin
et al. 1977). In a similar study by Manni et al. (1979), 128
of 156 breast cancer patients were determined to have a
complete hypophysectomy as shown by the fact that their
serum PRL was undetectable even after stimulation with
chlorpromazine. However, 28 patients (18%) had low to
normal levels of PRL which did not respond to
stimulation. This latter group were catagorized as
incomplete hypophysectomy, but no confirming data
were presented, leaving open the possibility that the PRL
was from another source which would not necessarily be

391

Vonderhaar: Prolactin and breast cancer


subject to the same stimulatory mechanisms. In addition,
low levels of circulating, bioactive PRL persisted in
patients given bromocriptine plus a somatostatin analog to
suppress pituitary hormone synthesis and activity
(Anderson et al. 1993). Prolactin is generated by tumors
as well as by a variety of normal tissues. Placenta is the
richest source of extrapituitary PRL (Sinha 1995). The
decidua is responsible for the high concentrations of PRL
in human amniotic fluid. The brain, uterus, dermal fibroblasts and the immune system all produce PRL (Clevenger
et al. 1990, Gellerson et al. 1994, Sinha 1995, Richards &
Hartman 1996). Several laboratories, including our own,
have shown that PRL is produced by both normal and
malignant mammary epithelial cells and thus may be an
autocrine/paracrine factor for this tissue. Steinmetz et al.
(1993) showed by in situ hybridization that PRL gene
transcripts are present in secretory mammary epithelial
cells from pregnant rats. Northern analysis and the reverse
transcriptase-polymerase chain reaction (RT-PCR) have
provided evidence for local synthesis of PRL by
mammary glands from lactating rats (Kurtz et al. 1993),
goats and sheep (LeProvost et al. 1994). More recently, the
presence of PRL mRNA has been demonstrated in human
breast cancer cells lines in our laboratory (Ginsburg &
Vonderhaar 1995) and in some primary human breast
carcinomas (Clevenger et al. 1995a).
In addition, we demonstrated that bioactive PRL is
synthesized by human breast cancer cells in culture and

acts in an autocrine manner to stimulate cell proliferation.


Growth of both T47Dco (ER-negative) and MCF-7 (ERpositive) human breast cancer cell lines was inhibited by
20-90% when cells were treated with monoclonal
antibodies raised against human pituitary PRL (Ginsburg
& Vonderhaar 1995). In addition, antisense RNA directed
against the gene encoding for pituitary PRL significantly
inhibited growth (>50%) of T47Dco cells (Ginsburg et al.
1997). Parallel cultures treated with a randomized
antisense RNA sequence grew at the same rate as
untreated controls. The presence of the mRNA for PRL in
T47Dco and MCF-7 cells was confirmed by RT-PCR,
followed by Southern analysis using pituitary hPRL
cDNA as the probe (Ginsburg & Vonderhaar 1995). In
total, 82% of all breast cancer cell lines tested (Table 1)
contained mRNA for PRL (Ginsburg et al. 1997).
Using 35 S-cysteine to metabolically label the T47Dco
cells, the nature of the protein synthesized was assessed.
Both cell extracts and conditioned media contain a 22 kDa
protein precipitated by anti-hPRL monoclonal antibodies.
Concentrated conditioned media prepared from T47Dco
cells stimulated the growth of PRL-responsive Nb2 rat
lymphoma cells to grow; these cells respond to picogram
quantities of lactogens. The level of biological activity in
the conditioned media is equivalent to 0.7 g/ml (14.5 pg
PRL/cell) pituitary PRL as measured by the Nb2 assay and
is approximately 30% of the amount normally produced
by the rat pituitary cell line, GH3 (Bancroft & Tashjian

Figure 1 The autocrine/paracrine action of PRL in the mammary gland.

392

Endocrine-Related Cancer (1999) 6 389-404


1971, Tanaka et al. 1980). By using a specific RIA for
human pituitary PRL, 0.35 g/ml PRL protein is detected.
When the media are pretreated with an anti-pituitary PRL
antibody, the activity in the conditioned media, like that of
the human pituitary PRL, is abolished (Ginsburg &
Vonderhaar 1995).
More than 80% of all breast cancer cell lines tested
contain mRNA for PRL (Fig. 2). In this sample of cell
lines, we found no correlation of ER status with the presence of PRL receptors or with the ability of the cells to
synthesize PRL (Table 1). PRL gene transcripts were also
found in tumors arising from MCF-7 cells implanted into
nude mice (Ginsburg et al. 1997). Immunologically detectable PRL was present in 60-85% of human breast cancer
biopsies, although the source of PRL is indeterminate
(Purnell et al. 1982, Agarwal et al. 1989). More than 75%
of primary breast cancer surgical samples also contain
mRNA for PRL. In the majority of cases, the amount of
mRNA for PRL and its receptors is significantly elevated
in cancerous vs the adjacent, non-involved tissue from the
same patient (Fields et al. 1993, Ginsburg et al. 1997).
Similar results were reported by Touraine et al. (1998)
who found PRL mRNA in all breast samples tested from
29 patients.
Post translational modifications of PRL
A PRL mRNA variant lacking exon 4 which arises from
alternate splicing has been reported in rat brain (Emanuele
et al. 1992). However, the mRNA for PRL isolated from
late pregnant and lactating sheep and goat mammary
glands differs from pituitary transcripts by only three mutations, two of which are silent (LeProvost et al. 1994).
Our recent data (E Ginsburg & BK Vonderhaar, unpublished observations) suggest at least 90% sequence identity between the PRL mRNA from the pituitary and from
human breast cancer cells. Similar results have been
reported in a variety of human breast cancer cell lines and
neoplastic breast tissue samples (Shaw-Bruha et al. 1997).
Because there is significant evidence that many diverse
activities of PRL are modulated by different post-translational modifications, the key to the role of autocrine PRL
in human breast cancer may lie therein. The biological
activity and immunoreactivity of pituitary PRL are modified by phosphorylation and glycosylation (Markoff et al.
1988, Lewis et al. 1989). Phosphorylated PRL, present in
murine, bovine and avian species (Walker 1994), has less
activity compared with the nonphosphorylated form
(Wicks & Brooks 1995). The phosphorylated form of PRL
is unable to bind to the receptor due to conformational
change in the hormone. The biological activity is restored
by dephosphorylation.
Glycosylation may selectively down-regulate PRL
action in target tissues (Hoffman et al. 1993). Glycosylated PRL, in the mammary casein synthesis and the Nb2

Figure 2 Presence of PRL transcripts in


human breast cancer cell lines and tumor from
nude mice. The PCR product obtained by
using primers designed from the known
sequence of pituitary PRL (Ginsburg &
Vonderhaar 1995) was electrophoresed on a
2% agarose gel and a predicted 612 bp PCR
product for PRL was detected by ethidium
bromide staining. Samples from human breast
cancer cell lines and tumor are marked as
indicated. M=600 bp marker. Reprinted from
Vonderhaar (1999).

proliferation assays, has biological activity similar to or


lower than that of the nonglycosylated form (Sinha et al.
1991, Pellegrini et al. 1992). Receptor binding activity and
immunological cross-reactivity are greatly reduced as a
result of glycosylation. Both glycosylated and nonglycosylated recombinant human PRL can be purified from the

393

Vonderhaar: Prolactin and breast cancer

murine C127 cell expression system. The 23 kDa nonglycosylated form of the PRL is 3-4 times more active in
the Nb2 mitogenesis bioassay compared with the 25 kDa
glycosylated form (Price et al. 1995). The changes in the
ratio of glycosylated and nonglycosylated forms of PRL
may explain physiologically diverse effects of PRL on
target tissues (Young et al. 1990).
The nature of the post-translational alteration of PRL
synthesized by the mammary gland and breast cancer cells
is unknown. Our observations that a panel of monoclonal
antibodies directed against pituitary PRL vary in their
ability to recognize the PRL produced by breast cancer
cells (Ginsburg & Vonderhaar 1995) suggests that there
are marked differences in post-translational modifications
between the product of the pituitary and that of the
mammary gland.
Cleaved prolactin and tumor angiogenesis
The mammary gland cleaves PRL (Wong et al. 1986);
three PRL species (25 kDa, 23 kDa and 14 kDa) are
detected in rat tissue. Following bromocriptine treatment,
the 25 kDa and 14 kDa forms remain, suggesting that they
are the products of the mammary gland and that cleavage
takes place (Lkhider et al. 1996). Explants of normal
mouse mammary tissue cleave PRL to yield two fragments (Baldocchi et al. 1995); the larger fragment of
cleaved PRL, either 14 kDa or 16 kDa, has anti-angiogenic
activity and inhibits the vascular endothelial growth factor
(VEGF)-induced growth of capillary endothelial cells
(Ferrara et al. 1991, DAngelo et al. 1995). VEGF is
essential for initial but not continued in vivo growth of
human breast carcinoma cells (Yoshiji et al. 1997). Explants from a transplantable rat tumor are unable to cleave
PRL (Baldocchi et al. 1995).
Regulation of expression of prolactin
The regulatory mechanisms which control PRL synthesis
and secretion by human breast cancer cells are not known.
In the normal sheep and goat mammary gland, the PRL
gene appears to be transcribed from the same proximal
promoter used by the pituitary (LeProvost et al. 1994).
However, other studies (Shaw-Bruha et al. 1997) suggest
that, in human breast cancer cells, PRL synthesis is regulated by the distal promoter used by decidua and lymphocytes. Our data (E Ginsburg & BK Vonderhaar, unpublished observations) show that one of the most effective
regulators of mammary PRL synthesis is the hor-mone
itself, suggesting an autoregulatory, feedback mechanism.
Such a mechanism may explain the obser-vation that treatment of lactating rats with bromocriptine results in a
decrease in the PRL localization to the endocytic organelles and an increase in localization to the organelles

394

associated with synthesis and exocytosis (Lkhider et al.


1996).

Prolactin receptors
PRL receptors belong to the cytokine hematopoietic
family of receptors (Kelly et al. 1991). The members of
this superfamily are single membrane-spanning receptors
organized into three domains comprising an extracellular
ligand binding domain, a hydrophobic transmembrane domain and an intracellular domain containing a proline-rich
motif. The three different forms of the PRL receptor differ
in their cytoplasmic domain. The long (90 kDa) and short
(40 kDa) forms of the receptor are generated by differential splicing of a single gene and differ only in the length
of the cytoplasmic domain (Kelly et al. 1993). The
intermediate form of the receptor is a deletion mutant of
the long form, lacking 198 amino acids in its cytoplasmic
region. This form of the receptor is found in Nb2 rat
lymphoma cells and is more sensitive to PRL compared
with the other forms (Ali et al. 1992). It may be present in
some human breast cancers (Clevenger et al. 1995a). Both
the long and the intermediate forms of the PRL receptor
transduce a differentiation signal as measured by induction of milk protein gene expression (Lesueur et al. 1991).
All three forms promote mitosis (ONeal & Yu Lee 1994,
Das & Vonderhaar 1995).
Both normal and malignant mammary cells contain
both long and short forms of the receptor. The receptors
also exist as multiple charged forms. Dimerization of the
membrane-associated receptor results from the interaction
of PRL with its receptor (Goffin & Kelly 1997, Sakal et al.
1997). The physiological significance of homodimers vs
heterodimers is unknown.
In human breast cancer, there is a clear correlation of
ER and progesterone receptor (PR) status with a variety of
disease parameters. However, because of the multiple size
and charged forms of the PRL receptor, no similar clear
correlation has been established for this hormone. PRL
receptors have been demonstrated in over 70% of breast
cancer biopsy samples using specific binding assays
(Codegone et al. 1981, Peyrat et al. 1981, Bonneterre et al.
1982, LHermite-Baleriaux et al. 1984). In our study
(Mertani et al. 1998), virtually all breast cancer surgical
samples were positive for PRL receptor mRNA by in situ
hybridization, but the amount varied considerably. Quantitative estimation of receptor mRNA levels was regionally
measured in areas corresponding to tumor cells or adipose
cells in the same section. No correlation emerged according to the histological type of lesion, probably due to the
large individual variation. Receptor immunoreactivity was
detected in some scattered stromal cells but the labeling
intensity was always weaker than for the neoplastic

Endocrine-Related Cancer (1999) 6 389-404


epithelial cells. In addition, the expression of mRNA as
measured by RT-PCR in malignant tissue was always
greater than in adjacent uninvolved tissue. Touraine et al.
(1998) made a similar observation in tissue from 29
patients using quantitative PCR and immunohisto-chemistry. In our study (Mertani et al. 1998) the expression of
PRL receptor occurred regardless of the ER or PR status.
In a similar study, Reynolds et al. (1997) demonstrated by
immunocytochemistry that >95% of breast cancers and
>93% of normal breast tissues expressed the PRL receptor.
They also found no association between the expression of
PRL receptor and ER or PR status. These observations are
in contrast to the report from Ormandy et al. (1997), who
found that the level of PRL receptor expression in breast
cancer cell lines was linearly related to that of the ER and
PR status.
In human breast cancers, the ratio of long and short
forms is unknown. The short form of the human PRL
receptor has not yet been cloned. No message for the
intermediate form of the receptor has been detected in our
samples, in contrast to the report by Clevenger et al.
(1995a).

Prolactin mitogenic signaling


pathways
PRL receptor-associated kinases
Because PRL receptors do not have intrinsic kinase
activity, they recruit several different kinases to transduce
the mitogenic signal. JAK2, a member of the Janus family
of kinases which is associated with the PRL receptor constitutively, is phosphorylated upon PRL binding in rat
lymphoma (David et al. 1994, Rui et al. 1994), mouse
mammary explants (Campbell et al. 1994) and murine
lymphoid BAF-3 cells (Dusanter-Fourt et al. 1994). In
murine lymphoid BAF-3 cells transfected with the long
form of the PRL receptor, JAK1 is associated with the
PRL receptor and is phosphorylated upon ligand binding
(Dusanter-Fourt et al. 1994). The JAK kinase is transphosphorylated and activated as a result of receptor dimerization. Phosphorylation of the JAK proteins may be one of
the earliest cellular events in response to the hormone
which ultimately triggers a chain of events in the PRL
signaling pathway.
The PRL receptor itself is phosphorylated within 1
min of PRL treatment, both in vivo in rabbit mammary
gland and in vitro in Chinese hamster ovary (CHO) cells
transfected with the long form of the PRL receptor cDNA
(Waters et al. 1995). Binding of PRL to its receptors then
induces phosphorylation of cytoplasmic transcription factors, mainly signal transducer and activator of transcription (STAT)-1, STAT-3 and STAT-5 (DaSilva et al.
1996). STAT-5 is activated for PRLs differentiation signal

(Wakao et al. 1994). Two different STAT-5 proteins have


been isolated from mouse mammary tissue (STAT-5a;
STAT-5b). Both of the transcription factors recognize
GAS sequences. Their expression is concurrent during
mammary gland development, increasing from the virgin
state, reaching a maximum at day 16 of pregnancy and
declining during lactation (Liu et al. 1995). STAT-5a is
essential for full lobuloalveolar development and lactation, as demonstrated by use of knockout mice (Liu et al.
1996). The phosphorylation of STAT proteins has also
been reported in Nb2 cells (David et al. 1994) and in
normal mammary epithelial cells (HC11) (Wakao et al.
1994). We have reported activation of STAT-5 upon PRL
treatment in T47D breast cancer cells (Das & Vonderhaar
1996b). DaSilva et al. (1996) also reported the phosphorylation of STAT-1, STAT-3 and STAT-5 in T47D cells
within 15 min of PRL treatment. Activation of STAT
proteins results in translocation of the transcription factors
to the nucleus and activation of gene transcription (Darnell
et al. 1994). STAT-1, and possibly STAT-3, is constitutively activated in breast cancer tissue (Watson & Miller
1995).
In hepatocytes of lactating rats in which the short form
of the receptor is predominant, PRL induces the association of PRL receptors with pp60 c-src and activation of its
tyrosine kinase activity (Berlanga et al. 1995). PRL stimulation in rat lymphoma Nb2 cells induces the association
and activation of src family protein tyrosine kinase p59
fyn (Clevenger & Medaglia 1994) and also the guanine
nucleotide releasing factor (GNRF)-vav (Clevenger et al.
1995b). In Nb2 cells, a protein tyrosine phosphatase,
PTP1D, which is associated with the PRL receptor-JAK2
complex is essential for PRL signal transduction during
induction of -casein (Ali et al. 1996). Both tyrosine
kinase inhibitors herbimysin A and tyrphos-tin are able to
decrease the expression of a -lactoglobin promoter/chloramphenicol acetyl transferase (CAT) con-struct by over
50% in CHO cells transfected with the rabbit PRL receptor. In addition, orthovanadate, an inhibitor of tyro-sine
phosphatase, is able to substitute for PRL in inducing CAT
responses in these cells (Daniel et al. 1996). This suggests
an intricate role of both kinases and phosphatases during
PRL signal transduction.
The ras-mitogen activated protein kinase
(MAPK) pathway
A variety of growth factors and cytokines mediate
proliferation by activating the ras-MAP kinase pathway of
signal transduction. By measuring the guanine nucleotides
bound to the protein, it is possible to demonstrate activation of ras p21 protein by PRL in a variety of cell systems
(Erwin et al. 1995, Elberg et al. 1996). Raf-1, mitogen
activated protein kinase kinase (MEK) and MAP kinase

395

Vonderhaar: Prolactin and breast cancer


are downstream kinases in the ras-MAPK pathway which
are activated for mitosis induced by PRL. In Nb2 cells,
PRL rapidly phosphorylates c-raf-1 (Clevenger et al.
1994) which is closely associated with the PRL receptor.
We have reported the rapid and transient activation of
these enzymes in both normal and breast cancer cells in
response to 5 min PRL treatment (Das & Vonderhaar
1996a). Tyrosine kinase inhibitors block the MAP kinase
activity and PRL-induced growth in these mammary cells
(Das & Vonderhaar 1996a). Both the long and the short
form of the PRL receptor are able to activate MAP kinase,
with the activity reaching a peak within 5 min of PRL
treatment (Das & Vonderhaar 1995). PRL also activates
MAP kinase in rat liver, in vivo, and in Nb2 cells (Buckley
et al. 1994, Piccoletti et al. 1994). Activation and nuclear
translocation of protein kinase C (PKC) and MAP kinase
have been reported during PRL-induced proliferation of
Nb2 cells (Buckley et al. 1994, Ganguli et al. 1996). A
preliminary report has shown PRL activation of 40S ribosomal protein S6 kinase (Carey & Liberti 1995), an
enzyme downstream from MAP kinase in this signal
cascade in Nb2 cells. The activation of S6 kinase reached
its peak at 1.5-2 h after PRL treatment. Rapamycin, a
specific inhibitor of S6 kinase, inhibited PRL-induced
mitogenesis (Carey & Liberti 1995).
The PKC pathway
The PKC pathway of signal transduction may be involved
in PRL-induced cell proliferation (Kiley et al. 1996). The
phosphoinositide cycle is activated during PRL signal
transduction, which then activates PKC. The phosphorylation and activation of phosphatidylinositol-3 (PI3)kinase for signal transduction of PRL in Nb2 cells have
been reported (Alsakkaf et al. 1996). Endogenously added
phospholipase C, an enzyme that hydrolyzes L--phosphoinositol 4,5-diphosphate (PIP2) to D-myo-inositol
triphosphate (IP3) and diacylglycerol, elicited PRL-like
effects on ornithine decarboxylase activity and RNA synthesis in pregnant mouse mammary gland explants in
culture (Rillema et al. 1983). In NOG-8 mammary cells,
PRL activates and translocates PKC to the membranes
within 5-10 min (Banerjee & Vonderhaar 1992). In addition, activation of PKC by PRL occurs in splenocytes,
liver, hypothalamus and Nb2 cells (Buckley et al. 1988,
Rillema et al. 1989). Waters and Rillema (1989) showed
translocation of PKC upon PRL treatment in explants
from mouse mammary glands. In addition activation of
nuclear PKC by PRL has been reported in Nb2 cells
(Buckley et al. 1988, Fan & Rillema 1993, Ganguli et al.
1996).
The SHC-Grb pathway
The JAK-SHC pathway is activated in 3T3-F442A cells
through the GH receptor which is a member of the same
396

receptor family as the PRL receptor (VanderKuur et al.


1995). The association of JAK2 with SHC occurs in a
rapid and transient manner in Nb2 cells upon 15 min of
PRL treatment (Erwin et al. 1995). In breast cells, both
normal and malignant, we have shown that PRL can
phosphorylate SHC proteins within 1 min of hormone
treatment followed by association with the Grb2-son of
sevenless (Sos) complex in these cells. Also, JAK2 is
rapidly phosphorylated and associated with the SHC
protein upon PRL binding to its receptors (Das &
Vonderhaar 1996b). Thus for PRL signal transduction
during the induction of the mitogenic response in human
breast cancer cells, both the JAK-STAT and the SHC-rasMAP kinase pathways are operative. Possible cross talk
between the MAP kinase pathway and the JAK-STAT
pathway has been suggested (David et al. 1995).
Early response genes
Cell growth results from modulation of early response
genes and late response genes in the signaling pathways of
a mitogen. Expression of a variety of early genes,
including myc (Zabala & Garcia-Ruiz 1989), is induced
by PRL in a number of systems. After 30 min PRL
treatment in hepatocytes, there is a rapid induction of the
expression of the proto-oncogenes c-fos, c-jun and c-src,
even in the presence of cycloheximide (Berlanga et al.
1995). This suggests that PRL stimulates the expression of
genes with AP-1 or serum response elements sequences in
their promoter regions. An early response gene, Pim-1,
which is a proto-oncogene encoding a conserved cytosolic
serine/threonine protein kinase, is stimulated by PRL
during mitogenesis of Nb2 and hematopoietic cells
(Buckley et al. 1995). Peak expression occurs at 2-4 h of
PRL treatment and is not affected by cycloheximide.
Interferon regulatory factor-1 (IRF-1), another early
activation gene, is induced over 20-fold in Nb2 cells by
PRL. This gene is induced twice by PRL in a single cell
cycle. First, it is induced during G1 at 30-60 min and then
again during early S phase at 10-12 h of hormone
treatment (Stevens et al. 1995). The second peak of IRF-1
mRNA expression in early S phase is dependent on the
continuous presence of PRL throughout G1, and is
correlated tightly with DNA synthesis and subsequent cell
proliferation. The GAS site in the IRF-1 promoter is
thought to act as a PRL responsive element that responds
to the mitogenic signal of PRL in T cells. Its activation in
breast cells has not been established.
Late response genes
Several cyclins, important regulators of cell cycle progression (Musgrove et al. 1996), are amplified in malignant
cells compared with their normal counterpart. Cyclin D1
is among the most commonly overexpressed oncogenes in
breast cancer. Cyclin D1 knockout mice are devoid of

Endocrine-Related Cancer (1999) 6 389-404


PRL-dependent lobuloalveolar structures in the mature
gland (Sicinski et al. 1995). In addition, following PRL
stimulation of quiescent Nb2 cells, the cyclin D2 mRNA
level increases in mid G1 phase and decreases sharply
before S phase. The cyclin D3 level increases in late G1/
early S phase and gradually decreases during S phase
(Hosokawa et al. 1994). Further elucidation of the
activation of specific late response genes should help
clarify the distinguishing events in PRL-induced mitogenesis vs differentiation.

Inhibition of prolactin action


Because the mammary gland synthesizes its own PRL and
the majority of human breast tumors contain PRL receptors, clinical approaches to controlling this disease in
PRL-responsive tumors need to incorporate antagonists of
PRL acting directly at the receptor level. Included among
these are mutants of hGH and hPRL, some of which have
been shown to have anti-prolactin activity in vitro under
defined growth conditions (Fuh et al. 1992, 1993, Fan &
Rillema 1993, Dattani et al. 1995, Fuh & Wells 1995,

Figure 3 Inhibition of ligand-induced receptor dimerization. (A) Sequential hormone


binding first through binding site 1, forming an inactive complex. The hormone then
binds to a second receptor through site 2, which leads to dimerization and formation
of an active complex. (B) Site 2 mutants prevent formation of active dimers. (C)
Inhibitors, such as TAM, interfere with the ability of the hormone to bind to the receptor.
Reprinted from Vonderhaar (1999).

397

Vonderhaar: Prolactin and breast cancer


Mode et al. 1996). These mutants usually involve that
portion of the hormone identified as site 2 (Fig. 3) where
amino acids with small side chain residues are replaced
with amino acids carrying large side chains. Steric hindrance prevents binding of the mutant to the second
receptor in the dimer and hence they act as hormone
antagonists (Goffin & Kelly 1997).
The anti-lactogen binding site (ALBS)
In addition, tamoxifen (TAM), the first line of therapy in
pre- and post-menopausal, ER-positive breast cancer
patients, is also an anti-prolactin or anti-lactogen
(Vonderhaar & Banerjee 1991). In some reports, TAM has
been shown to be effective in a small number of ERnegative breast cancer patients (Hawkins et al. 1980, Furr
& Jordan 1984, Nolvadex Adjuvant Trial Organization
1985). Work in vitro has also suggested that TAM can
work through other ER-independent mechanisms. Frequently, estrogens can reverse TAM inhibition of cell
proliferation in vitro when the anti-estrogen is used at low
concentrations. However, the growth rates cannot be
restored by estradiol in the presence of higher concentrations of the anti-estrogen (Sutherland et al. 1987).
Growth of several ER-negative human breast cancer cell
lines is inhibited by micromolar concentrations of TAM
(Green et al. 1981, Chouvet et al. 1988, Das et al. 1994,
Perry et al. 1995). Several other cellular proteins besides
the ER may be directly affected by TAM (Vonderhaar &
Banerjee 1991). Pollak et al. (1990) have suggested that,
during TAM therapy, there is a decrease in circulating

IGF-I which could account for the ER-independent response in some patients. Others have suggested that direct
interaction of TAM with PKC may be responsible
(OBrian et al. 1985). We have shown that the antilactogenic activity of TAM results from interaction with
the ALBS (Das et al. 1993) which is located on the PRL
receptor.
The ALBS is a member of the family of high affinity
membrane-associated binding sites called anti-estrogen
binding sites or AEBS (Sutherland & Foo 1979,
Sutherland et al. 1980) which have been identified in a
variety of tissues (Gulino & Pasqualini 1982, Mehta et al.
1984). A TAM resistant clone of MCF-7 cells, RTx6,
differs from the parent MCF-7 cells in having no AEBS,
but is identical to the parent cells with respect to the ER
content and affinity (Faye et al. 1987). These cells could
not be inhibited by TAM, whereas MCF-7 cells were.
Anti-estrogens, acting through the ALBS, inhibit the
growth of PRL-responsive cells, even in the absence of ER
(Das et al. 1994, Das & Vonderhaar 1997a). The ALBS,
located on cellular membranes, binds TAM and related
non-steroidal anti-estrogens with high affinity, but does
not bind estrogens (Fig. 4) (Biswas & Vonderhaar 1991,
Vonderhaar & Banerjee 1991). It is through the ALBS that
TAM inhibits PRL-induced growth of ER-negative, Nb2
rat lymphoma cells (Biswas & Vonderhaar 1989); the
effects are not reversed by estradiol. In addition, binding
of PRL to particulate and solubilized microsomal membranes isolated from normal mammary glands of lactating
mice is inhibited by direct addition of 10-10 M or greater

Figure 4 Interactions at the anti-lactogen binding site (ALBS) on the PRL receptor.
E2, estradiol. Reprinted from Vonderhaar (1999).

398

Endocrine-Related Cancer (1999) 6 389-404

concentrations of TAM to the binding assays (Biswas &


Vonderhaar 1991); estradiol did not have this effect. Maximal inhibition of PRL binding by TAM is observed in the
light microsomes that contain the plasma membranes.
Characterization of the relationship of the ALBS to the
PRL receptor (Das et al. 1993), showed that TAM decreases the number of binding sites without changing the
receptors affinity for PRL. TAM acts by inhibiting the
binding of PRL to its receptor (Fig. 3), rather than promoting dissociation of the hormone-receptor complex. The
PRL receptor and the ALBS co-purify on either PRLSepharose or TAM-Sepharose affinity columns. Both are
recognized by the anti-PRL receptor monoclonal antibody
B6.2 (Das et al. 1993). ER-negative T47Dco as well as
ER-positive T47D cells bind lactogenic hormones specifically and this binding is inhibited by 30-50% when 10-11
M TAM is added directly to the whole cell binding reaction (Das et al. 1994). At 10-9 M, TAM inhibits PRL
binding by 70-90% with complete inhibition at 10-7 M
TAM. Subsequently, PRL-induced growth for both cell
lines is inhibited. In the ER-negative mouse mammary
epithelial cell line, NOG-8, TAM inhibits PRL binding
and at concentrations as low as 10-9 M rapidly inhibits
PRL signal transduction through raf-1, MEK and MAP
kinase. This, in turn, leads to inhibition of PRL-induced
growth of these PRL-responsive mammary cells (Das &
Vonderhaar 1997b). Thus, we conclude that the ALBS is
on the PRL receptor, and that TAM and other related, nonsteroidal, triphenylethylene anti-estrogens may inhibit
growth of ER-negative human breast cancer cells in vitro
through this mechanism.

Conclusion
The ability of PRL to stimulate growth of human breast
cancer cells in culture, coupled with the presence of active
receptors for PRL on the majority of breast carcinomas,
suggests that this peptide hormone is an active player in
this disease. Understanding its role is complicated by the
fact that human breast cancer cells synthesize and secrete
significant amounts of biologically active PRL. These data
suggest that clinically useful reagents should be sought
which act at the level of the target tissue. The drug tamoxifen, in addition to its action as an anti-estrogen, is also an
anti-lactogen and hence may be a useful tool for understanding the potential mechanism of action of PRL in
tumors which are PRL receptor-positive even if they are
ER-negative.

References
Agarwal PK, Tandon S, Agarwal AK & Kumar S 1989 Highly
specific sites of prolactin binding in benign and malignant
breast disease. Indian Journal of Experimental Biology 27
1035-1038.

Ali S, Edery M, Pellegrini I, Lesueur L, Paly J, Djiane J & Kelly


PA 1992 The Nb2 form of prolactin receptor is able to activate
a milk protein gene promoter. Molecular Endocrinology 6
1242-1248.
Ali S, Chen Z, Lebrun J-J, Vogel W, Kharitonenkov A, Kelly P
& Ullrich A 1996 PTP1D is a positive regulator of the
prolactin signal leading to -casein promoter activation.
EMBO Journal 15 135-142.
Alsakkaf KA, Dobson PRM & Brown BL 1996 Activation of
phosphatidylinositol 3-kinase by prolactin in Nb2 cells.
Biochemical and Biophysical Research Communications 221
779-784.
Anderson E, Ferguson JE, Morten H, Shalet SM, Robinson EL &
Howell A 1993 Serum immunoreactive and bioactive
lactogenic hormones in advanced breast cancer patients
treated with bromocriptine and octreotide. European Journal
of Cancer 29A 209-217.
Baldocchi RA, Tan L, Hom YK & Nicoll CS 1995 Comparison
of the ability of normal mouse mammary tissues and
mammary adenocarcinoma to cleave rat prolactin.
Proceedings of the Society for Experimental Biology and
Medicine 208 283-287.
Bancroft FC & Tashjian AH 1971 Growth in suspension culture
of rat pituitary cells which produce growth hormone and
prolactin. Experimental Cell Research 64 125-128.
Banerjee R & Vonderhaar BK 1992 Prolactin induced protein
kinase C activity in a mouse mammary epithelial cell line
NOG-8. Molecular and Cellular Endocrinology 90 61-67.
Barni S, Lissoni P, Brivio F, Fumagalli L, Merlini D, Cataldo M,
Rovelli F & Tancini G 1994 Serum levels of insulin-like
growth factor-I in operable breast cancer in relation to the
main prognostic variables and their perioperative changes in
relation to those of prolactin. Tumori 80 212-215.
Berlanga JJ, Vara JAF, Martin-Perez J & Garcia-Ruiz JP 1995
Prolactin receptor is associated with c-src kinase in rat liver.
Molecular Endocrinology 9 1461-1467.
Bhatavdekar JM, Patel DD, Vora HH, Ghosh N, Shah NG,
Karelia NH, Suthar TP, Dave RS & Balar DB 1994 Nodepositive breast cancer: prognostic significance of the plasma
prolactin compared with steroid receptors and clinicopathological features. Oncology Reports 1 841-845.
Biswas R & Vonderhaar BK 1987 Role of serum in prolactin
responsiveness of MCF-7 human breast cancer cells in long
term tissue culture. Cancer Research 47 3509-3514.
Biswas R & Vonderhaar BK 1989 Antiestrogen inhibition of
prolactin induced growth of the Nb2 rat lymphoma cell line.
Cancer Research 49 6295-6299.
Biswas R & Vonderhaar BK 1991 Tamoxifen inhibition of
prolactin action in the mouse mammary gland. Endocrinology
128 532-538.
Bonneterre J, Peyrat JP, Vandewalle B, Beuscart R, Vie MC &
Cappelaere P 1982 Prolactin receptors in human breast cancer.
European Journal of Cancer and Clinical Oncology 18 11571162.
Boot LM, Muhlbock O & Ropcke G 1962 Prolactin and the
induction of mammary tumors in mice. General and
Comparative Endocrinology 2 601-602.

399

Vonderhaar: Prolactin and breast cancer


Boyns AR, Buchan R, Cole EN, Forrest APM & Griffiths K 1973
Basal prolactin blood levels in three strains of rat with
differing incidence of 7,12-dimethylbenzanthracene-induced
mammary tumors in rats. European Journal of Cancer 9 169171.
Buckley AR, Crowe PD & Russell DH 1988 Rapid activation of
protein kinase C in isolated rat liver nuclei by prolactin, a
known hepatic mitogen. Proceedings of the National Academy
of Sciences of the USA 85 8649-8653.
Buckley AR, Rao YP, Buckley DJ & Gout PW 1994 Prolactininduced phosphorylation and nuclear translocation of MAP
kinase in Nb2 lymphoma cells. Biochemical and Biophysical
Research Communications 204 1158-1164.
Buckley AR, Buckley DJ, Leff MA, Hoover DS & Magnuson NS
1995 Rapid induction of pim-1 expression by prolactin and
interleukin-2 in rat Nb2 lymphoma cells. Endocrinology 136
5252-5259.
Burke RE & Gaffney EV 1978 Prolactin can stimulate general
protein synthesis in human breast cancer cells (MCF-7) in
long-term culture. Life Sciences 23 901-906.
Calaf G, Garrido F, Moyano C & Rodriguez R 1986 Influence of
hormones on DNA synthesis of breast tumors in culture.
Breast Cancer Research and Treatment 8 223-232.
Campbell GS, Argetsinger LS, Ihle JN, Kelly PA, Rillema JA &
Carter-Su C 1994 Activation of JAK (JAK2) tyrosine kinase
by prolactin receptors in NB (NB2) cells and mouse mammary
gland explants. Proceedings of the National Academy of
Sciences of the USA 91 5232-5236.
Carey GB & Liberti JP 1995 Stimulation of receptor-associated
kinase, tyrosine kinase, and MAP kinase is required for
prolactin-mediated macromolecular biosynthesis and mitogenesis in Nb2 lymphoma. Archives of Biochemistry and
Biophysics 316 179-189.
Chouvet C, Vicard E, Frappart L, Falette N, Lefebvre MF & Saez
S 1988 Growth inhibitory effect of 4-hydroxy-tamoxifen on
the BT-20 mammary cancer cell line. Journal of Steroid
Biochemistry 31 655-663.
Clevenger CV & Medaglia MV 1994 The protein tyrosine kinase
p59fyn is associated with prolactin (PRL) receptor and is
activated by PRL stimulation of T-lymphocytes. Molecular
Endocrinology 8 674-681.
Clevenger CV, Russell DH, Appasamy PM & Prystowsky MB
1990 Regulation of IL-2-driven T-lymphocyte proliferation
by prolactin. Proceedings of the National Academy of
Sciences of the USA 87 6460-6464.
Clevenger CV, Torigoe T & Reed JC 1994 Prolactin induces
rapid phosphorylation and activation of prolactin receptorassociated RAF-1 kinase in a T-cell line. Journal of Biological
Chemistry 269 5559-5565.
Clevenger CV, Chang WP, Ngo W, Pasha TLM, Montone KT &
Tomaszewski JE 1995a Expression of prolactin and prolactin
receptor in human breast carcinoma. American Journal of
Pathology 146 695-705.
Clevenger CV, Ngo W, Sokol DL, Luger SM & Gewirtz AM
1995b Vav is necessary for prolactin-stimulated proliferation
and is translocated into the nucleus of a T-cell line. Journal of
Biological Chemistry 270 13246-13253.

400

Codegone ML, DiCarlo R, Muccioli G & Bussolati G 1981


Histology and cytometrics in human breast cancers assayed
for the presence of prolactin receptors. Tumori 67 549-552.
DAngelo G, Struman I, Martial J & Weiner RI 1995 Activation
of mitogen-activated protein kinases by vascular endothelial
growth factor and basic fibroblast growth factor in capillary
endothelial cells is inhibited by the antiangiogenic factor 16kDa N-terminal fragment of prolactin. Proceedings of the
National Academy of Sciences of the USA 92 6374-6378.
Daniel N, Waters MJ, Bignon C & Djiane J 1996 Involvement of
a subset of tyrosine kinases and phosphatases in regulation of
the -lactoglobulin gene promoter by prolactin. Molecular
and Cellular Endocrinology 118 25-35.
Darnell Jr JE, Kerr IM & Stark GR 1994 Jak-STAT pathways and
transcriptional activation in response to IFNs and other
extracellular signaling proteins. Science 264 1415-1421.
Das R & Vonderhaar BK 1995 Transduction of prolactins
growth signal through both the long and short forms of the
prolactin receptor. Molecular Endocrinology 9 1750-1759.
Das R & Vonderhaar BK 1996a Activation of raf-1, MEK and
MAP kinase in prolactin responsive mammary cells. Breast
Cancer Research and Treatment 40 141-149.
Das R & Vonderhaar BK 1996b Involvement of SHC, Grb2, Sos
and ras in prolactin signal transduction in mammary cells.
Oncogene 13 1139-1145.
Das R & Vonderhaar BK 1997a Prolactin as a mitogen in
mammary cells. Journal of Mammary Gland Biology
Neoplasia 2 29-39.
Das R & Vonderhaar BK 1997b Tamoxifen inhibits prolactin
signal transduction in estrogen receptor negative NOG-8
mammary epithelial cells. Cancer Letters 116 41-46.
Das R, Biswas R & Vonderhaar BK 1993 Characteristics of the
antilactogen binding site in mammary gland membranes.
Molecular and Cellular Endocrinology 98 1-8.
Das R, Ginsburg E & Vonderhaar BK 1994 Tamoxifen as an
antilactogen in human breast cancer cells. In Proceedings of
the International Cancer Congress, pp 1487-1491. Eds RS
Rao, MG Deo, LD Sanghvi & I Mittra. Bologna: Monduzzi
Editore, Inc.
DaSilva L, Rui H, Erwin RA, Howard OMZ, Kirken RA,
Malabarba MG, Hackett RH, Larner A & Farrar WL 1996
Prolactin recruits STAT1, STAT3 and STAT5 independent of
conserved receptor tyrosine TYR402, TYR479, TYR515 and
TYR580. Molecular and Cellular Endocrinology 117 131140.
Dattani MT, Hindmarsh PC, Brook GD, Robinson ICAF,
Kopchick JJ & Marshall NJ 1995 G120R, a human growth
hormone antagonist, shows zinc-dependent agonist and
antagonist activity on Nb2 cells. Journal of Biological
Chemistry 270 9222-9226.
David M, Petricoin EF, Igarashi KI, Feldman GM, Finbloom DS
& Larner AC 1994 Prolactin activates the interferon-regulated
p91 transcription factor and the JAK2 kinase by tyrosine
phosphorylation. Proceedings of the National Academy of
Sciences of the USA 91 7174-7178.
David M, Petricoin III E, Benjamin C, Pine R, Weber MJ &
Larner AC 1995 Requirement for MAP kinase (ERK2)

Endocrine-Related Cancer (1999) 6 389-404


activity in interferon - and interferon -stimulated gene
expression through STAT proteins. Science 269 1721-1723.
Dusanter-Fourt I, Muller O, Ziemiecki A, Mayeux P, Drucker B,
Djiane J, Wilks A, Harpur AG, Fischer S & Gisselbrecht S
1994 Identification of JAK protein tyrosine kinases as
signaling molecules for prolactin. Functional analysis of
prolactin receptor and prolactin-erythropoietin receptor
chimera expressed in lymphoid cells. EMBO Journal 13 25382591.
Elberg G, Rapoport MJ, Vashdi-Elberg D, Gertler A & Scechter
Y 1996 Lactogenic hormones rapidly activate p21 ras/
mitogen-activated protein kinase in Nb2-11C rat lymphoma
cells. Endocrine 4 65-71.
Emanuele NV, Jurgens JK, Halloran MM, Tentler JJ, Lawrence
AM & Kelley MR 1992 The rat prolactin gene is expressed in
brain tissue: detection of normal and alternatively spliced
prolactin messenger RNA. Molecular Endocrinology 6 35-42.
Erwin RA, Kirken RA, Malabarba MG, Farrar WL & Rui H 1995
Prolactin activates Ras via signaling proteins SHC, growth
factor receptor bound 2 and son of sevenless. Endocrinology
136 3512-3518.
Fan G & Rillema JA 1993 Prolactin stimulation of protein kinase
C in isolated mouse mammary gland nuclei. Hormone and
Metabolic Research 25 564-568.
Faye JC, Fargin A, Valette A & Bayard F 1987 Antiestrogens,
different sites of action than the estrogen receptor? Hormone
Research 28 202-211.
Ferrara N, Clapp C & Weiner R 1991 The 16K fragment of
prolactin specifically inhibits basal and fibroblast growth
factor stimulated growth of capillary endothelial cells.
Endocrinology 129 896-900.
Fields K, Kulig E & Lloyd RV 1993 Detection of prolactin
messenger RNA in mammary and other normal and neoplastic
tissue by polymerase chain reaction. Laboratory Investigation
68 354-360.
Fuh G & Wells JA 1995 Prolactin receptor antagonists that inhibit
the growth of breast cancer cell lines. Journal of Biological
Chemistry 270 13133-13137.
Fuh G, Cunningham BC, Fukunaga R, Nagata S, Goeddel DV &
Wells JA 1992 Rational design of potent antagonists to the
human growth hormone receptor. Science 256 1677-1679.
Fuh G, Colosi P, Wood WI & Wells JA 1993 Mechanism-based
design of prolactin receptor antagonists. Journal of Biological
Chemistry 268 5376-5381.
Furr BJA & Jordan VC 1984 The pharmacology and clinical uses
of tamoxifen. Pharmacology and Therapeutics 25 127-205.
Ganguli S, Hu L, Menke P, Collier RJ & Gertler A 1996 Nuclear
accumulation of multiple protein kinases during prolactininduced proliferation of Nb2 rat lymphoma cells. Journal of
Cellular Physiology 167 251-260.
Gellerson B, Kempf R, Teglmann R & DiMattia GE 1994
Nonpituitary human prolactin gene transcription is
independent of pit-1 and differentially controlled in
lymphocytes and in endometrial stroma. Molecular
Endocrinology 8 356-373.
Ginsburg E & Vonderhaar BK 1995 Prolactin synthesis and
secretion by human breast cancer cells. Cancer Research 55
2591-2595.

Ginsburg E, Das R & Vonderhaar BK 1997 Prolactin: an


autocrine growth factor in the mammary gland. In Biological
Signalling in the Mammary Gland, pp 47-58. Eds CJ Wilde,
M Peaker & E Taylor. Ayr, Scotland: Hannah Research
Institute.
Goffin V & Kelly PA 1997 The prolactin/growth hormone
receptor family: structure/function relations. Journal of
Mammary Gland Biology Neoplasia 2 7-17.
Green MD, Whybourne AM, Taylor IW & Sutherland RL 1981
Effects of antioestrogens on the growth and cell cycle kinetics
of cultured human mammary carcinoma cells. In NonSteroidal Antiestrogens: Molecular Pharmacology and
Antitumor Actions, pp 397-412. Eds RL Sutherland & VC
Jordan. New York: Academic Press, Inc.
Gulino A & Pasqualini JR 1982 Heterogeneity of binding sites
for tamoxifen and tamoxifen derivatives in estrogen target and
non-target fetal organs of guinea pig. Cancer Research 42
1913-1921.
Haus E, Lakatua DJ, Halberg F, Halberg E, Cornelissen G,
Sackett LL, Berg HG, Kawasaki T, Ueno M, Uezono K,
Matsuoka M & Omae T 1980 Chronobiological studies of
plasma prolactin in women in Kyushu, Japan and Minnesota,
USA. Journal of Clinical Endocrinology and Metabolism 51
632-640.
Hawkins RA, Roberts MM & Forest APM 1980 Estrogen
receptors and breast cancer: current status. British Journal of
Surgery 67 153-169.
Henson JC, Coune A & Staquet M 1972 Clinical trial of 2-Br-aergocryptine (CB154) in advanced breast cancer. European
Journal of Cancer 8 155-156.
Hoffman T, Penel C & Ronin C 1993 Glycosylation of human
prolactin regulates hormone bioactivity and metabolic
clearance. Journal of Endocrinological Investigation 16 807816.
Holdaway IM, Mason BH, Gibbs EE, Rajasoorya C, Lethaby A,
Hopkins KD, Evans MC, Lim T & Schooler B 1997 Seasonal
variation in the secretion of mammotrophic hormones in
normal women and women with previous breast cancer.
Breast Cancer Research and Treatment 42 15-22.
Holtkamp W, Nagel GA, Wander HE, Rauschecker HF &
VonHeyden D 1984 Hyperprolactinemia is an indicator of
progressive disease and poor prognosis in advanced breast
cancer. International Journal of Cancer 34 323-328.
Hosokawa Y, Onga T & Nakashima K 1994 Induction of D2 and
D3 cyclin-encoding genes during promotion of the G1/S
transition by prolactin in rat Nb2 cells. Gene 147 249-252.
Ingram DM, Nottage EM & Roberts AN 1990 Prolactin and
breast cancer risk. Medical Journal of Australia 153 469-473.
Kelly PA, Djiane J, Postel-Vinay MC & Edery M 1991 The
prolactin/growth hormone receptor family. Endocrine Review
12 235-251.
Kelly PA, Ali S, Rozakis M, Goujon L, Nagano M, Pellegrini I,
Gould D, Djiane J, Edery M, Finidori J & Postel-Vinay MC
1993 The growth hormone/prolactin receptor family. Recent
Progress in Hormone Research 48 123-164.
Kiley SC, Welsh J, Narvaez CJ & Jaken S 1996 Protein kinase C
isozymes and substrates in mammary carcinogenesis. Journal
of Mammary Gland Biology Neoplasia 1 177-187.

401

Vonderhaar: Prolactin and breast cancer

Kurtz A, Bristol LA, Toth BE, Lazar-Wesley E, Takacs L &


Kacsoh B 1993 Mammary epithelial cells of lactating rats
express prolactin messenger ribonucleic acid. Biology of
Reproduction 48 1095-1103.
Lachelin GCL, Yen SSC & Alksne JF 1977 Hormonal changes
following hypophysectomy in humans. Obstetrics and
Gynecology 50 333-339.
Lemus-Wilson A, Kelly PA & Blask DE 1995 Melatonin blocks
the stimulatory effects of prolactin on human breast cancer
cell growth in culture. British Journal of Cancer 72 14351440.
LeProvost F, Leroux C, Martin P, Gaye P & Djiane J 1994
Prolactin gene expression in ovine and caprine mammary
gland. Neuroendocrinology 60 305-313.
Lesueur L, Edery M, Ali S, Paly J, Kelly PA & Djiane J 1991
Comparison of long and short forms of the prolactin receptor
on prolactin-induced milk protein gene transcription.
Proceedings of the National Academy of Sciences of the USA
88 824-828.
Lewis UL, Singh RNP & Lewis LJ 1989 Two forms of glycosylated prolactin have different pigeon crop sac-stimulating
activities. Endocrinology 124 1558-1563.
LHermite-Baleriaux M, Casteels S, Vokaer A, Loriaus C, Noel
G & LHermite M 1984 Prolactin and prolactin receptors in
human breast disease. Progress in Cancer Research and
Therapy 31 325-334.
Lissoni P, Barni S, Cazzaniga M, Ardizzoia A, Rovelli F, Tancici
G, Brivio F & Frigerio F 1995 Prediction of recurrence of
inoperable breast cancer by postoperative changes in prolactin
secretion. Oncology 52 439-442.
Liu X, Robinson GW, Gouilleux F, Groner B & Hennighausen L
1995 Cloning and expression of Stat5 and an additional
homologue (Stat5b) involved in prolactin signal transduction
in mouse mammary tissue. Proceedinsg of the National
Academy of Sciences of the USA 92 8831-8835.
Liu X, Robinson GW, Wagner KU, Garrett L, Wynshaw-Boris A
& Hennighausen L 1996 Stat5a is mandatory for adult mammary gland development and lactogenesis. Genes and
Development 11 179-186.
Lkhider M, Delpal S & Olivier-Bousquet M 1996 Rat prolactin
in serum, milk and mammary tissue: characterization and
intracellular localization. Endocrinology 137 4969-4979.
Love RR & Rose DP 1985 Elevated bioactive prolactin in women
at risk for familial breast cancer. European Journal of Cancer
and Clinical Oncology 21 1553-1554.
Love RR, Rose DR, Surawicz TS & Newcomb PA 1991 Prolactin
and growth hormone levels in premenopausal women with
breast cancer and healthy women with a strong family history
of breast cancer. Cancer 68 1401-1405.
McManus MJ & Welsch CW 1984 The effect of estrogen,
progesterone, thyroxine, and human placental lactogen on
DNA synthesis of human breast ductal epithelium maintained
in athymic nude mice. Cancer 54 1920-1927.
Malarkey WB, Kennedy M, Allred LE & Milo G 1983
Physiological concentrations of prolactin can promote the

402

growth of human breast tumor cells in culture. Journal of


Clinical Endocrinology and Metabolism 56 673-677.
Manni A, Pearson OH, Brodkey J & Marshall JS 1979 Transsphenoidal hypophysectomy in breast cancer: evidence for an
individual role of pituitary and gonadal hormones in
supporting tumor growth. Cancer 44 2330-2337.
Manni A, Wright C, Davis G, Glenn J, Joehl R & Feil P 1986
Promotion by prolactin of the growth of human breast
neoplasms cultured in vitro in the soft agar clonogenic assay.
Cancer Research 46 1669-1672.
Manni A, Boucher AE, Demers LM, Harvey HA, Lipton A,
Simmonds MA & Bartholomew M 1989 Endocrine effects of
combined somatostatin analog and bromocriptine therapy in
women with advanced breast cancer. Breast Cancer Research
and Treatment 14 289-298.
Markoff E, Sigel MB, Lacour N, Seavey BK, Friesen HG &
Lewis UJ 1988 Glycosylation selectively alters the biological
activity of prolactin. Endocrinology 123 1303-1306.
Mehta RG, Cerny WL & Moon RC 1984 Distribution of
antiestrogen-specific binding sites in normal and neoplastic
mammary gland. Oncology 41 387-392.
Mershon J, Sall W, Mitchner N & Ben-Jonathan N 1995 Prolactin
is a local growth factor in rat mammary tumors.
Endocrinology 136 3619-3623.
Mertani HC, Garcia-Caballero T, Lambert A, Gerard F, Palayer
C, Boutin JM, Vonderhaar BK, Waters MJ, Losie PE & Morel
G 1998 Cellular expression of growth hormone and prolactin
receptors in human breast disorders. International Journal of
Cancer 79 201-211.
Mode A, Tollet P, Wells T, Carmignac DF, Clark RG, Chen WY,
Kopchick JJ & Robinson ICAF 1996 The human growth
hormone (hGH) antagonist G120RhGH does not antagonize
GH in the rat, but has paradoxical agonist activity, probably
via the prolactin receptor. Endocrinology 137 447-454.
Muhlbock O & Boot LM 1959 Induction of mammary cancer in
mice without the mammary tumor agent by isografts of
hypophyses. Cancer Research 19 402-412.
Musgrove EA, Hui R, Sweeney KJE, Watts CKW & Sutherland
RL 1996 Cyclins and breast cancer. Journal of Mammary
Gland Biology Neoplasia 1 153-162.
Nagasawa H, Miur K, Niki K & Namiki H 1985 Interrelationship
between prolactin and progesterone in normal mammary
gland growth in SHN virgin mice. Experimental and Clinical
Endocrinology 86 357-360.
Nolvadex Adjuvant Trial Organization 1985 Controlled trial of
tamoxifen as single adjuvant agent in management of early
breast cancer. Lancet i 836-839.
OBrian CA, Liskamp RM, Solomon DH & Weinstein IB 1985
Inhibition of protein kinase C by tamoxifen. Cancer Research
45 2462-2465.
ONeal KD & Yu Lee L 1994 Differential signal transduction of
the short, Nb2, and long prolactin receptors. Journal of
Biological Chemistry 269 26076-26082.
Ormandy CJ, Hall RE, Manning DL, Robertson JFR, Blamey
RW, Kelly PA, Nicholson RI & Sutherland RL 1997
Coexpression and cross-regulation of the prolactin receptor

Endocrine-Related Cancer (1999) 6 389-404


and sex steroid hormone receptors in breast cancer. Journal of
Clinical Endocrinology and Metabolism 82 3692-3699.
Patel DD, Bhatavdekar JM, Chikhlikar PR, Ghosh N, Suthar TP,
Shah NG, Mehta RH & Balar DB 1996 Node negative breast
carcinoma: hyperprolactinemia and/or overexpression of p53
as an independent predictor of poor prognosis compared with
newer and established prognosticators. Journal of Surgical
Oncology 62 86-92.
Pearson OH & Manni A 1978 Hormonal control of breast cancer
growth in women and rats. In Current Topics in Experimental
Endocrinology, pp 75-92. Eds L Martini & VHT James. New
York: Academic Press.
Pellegrini I, Lebrun JJ, Ali S & Kelly PA 1992 Expression of
prolactin and its receptors in human lymphoid cells.
Molecular Endocrinology 6 1023-1031.
Perry RR, Kang Y & Greaves B 1995 Effects of tamoxifen on
growth and apoptosis of estrogen-dependent and -independent
human breast cancer cells. Annals of Surgical Oncology 2 238245.
Peyrat JP, DeWailly D, Djiane J, Kelly PA, Vandewalle B,
Bonneterre J & LeFebvre J 1981 Total prolactin binding sites
in human breast cancer biopsies. Breast Cancer Research and
Treatment 1 369-373.
Peyrat JP, Djiane J, Bonneterre J, Vandewalle B, Vennin P,
Delobelle A, Depadt G & Lefebvre J 1984 Stimulation of
DNA synthesis by prolactin in human breast tumor explants.
Relation to prolactin receptors. Anticancer Research 4 257261.
Piccoletti R, Maroni P, Bendinelli P & Bernelli-Zazzera A 1994
Rapid stimulation of mitogen-activated protein kinase of rat
liver by prolactin. Biochemical Journal 303 429-433.
Pollak M, Constantino J, Polychronakas C, Blauer HG, Guyda H,
Redmond C, Fisher B & Margolese R 1990 Effect of
tamoxifen on serum insulin-like growth factor-I levels of stage
I breast cancer patients. Journal of the National Cancer
Institute 82 1693-1697.
Price AE, Loginenko KB, Higgins EA, Cole ES & Richards S
1995 Studies on the microheterogeneity and in vitro activity
of glycosylated and nonglycosylated recombinant human
prolactin separated using a novel purification process.
Endocrinology 136 4827-4833.
Purnell DM, Hillman EA, Heatfield BM & Trump BF 1982
Immunoreactive prolactin in epithelial cells of normal and
cancerous human breast and prostate detected by the
unlabeled antibody peroxidase-antiperoxidase method.
Cancer Research 42 2317-2324.
Reynolds C, Montone KT, Powell CM, Tomaszewski JE &
Clevenger CV 1997 Expression of prolactin and its receptor in
human breast carcinoma. Endocrinology 138 5555-5560.
Richards RG & Hartman SM 1996 Human dermal fibroblast cells
express prolactin in vitro. Journal of Investigative
Dermatology 106 1250-1255.
Rillema JA, Wing LY & Foley KA 1983 Effects of
phospholipases on ornithine decarboxylase activity in
mammary gland explants from midpregnancy mice.
Endocrinology 113 2024-2028.
Rillema JA, Waters SB & Tarrant TM 1989 Studies on the
possible role of protein kinase C in the prolactin regulation of

cell replication in Nb2 node lymphoma cells. Proceedings of


the Society for Experimental Biology and Medicine 192 140144.
Rui H, Kirken RA & Farrar WL 1994 Activation of receptorassociated tyrosine kinase JAK2 by prolactin. Journal of
Biological Chemistry 269 5364-5368.
Sakal E, Elberg G & Gertler A 1997 Direct evidence that
lactogenic hormones induce homodimerization of membraneanchored prolactin receptor in intact Nb2-11C rat lymphoma
cells. FEBS Letters 410 289-292.
Salih H, Brander W, Flax H & Hobbs JR 1972 Prolactin
dependence in human breast cancers. Lancet ii 1103-1105.
Shafie SM & Grantham FH 1981 Role of hormones in the growth
and regression of human breast cancer cells (MCF-7)
transplanted into athymic nude mice. Journal of the National
Cancer Institute 67 51-56.
Shaw-Bruha CM, Pirrucello SJ & Shull JD 1997 Expression of
the prolactin gene in normal and neoplastic human breast
tissues and human mammary cell lines: promoter usage and
alternative mRNA splicing. Breast Cancer Research and
Treatment 4 243-253.
Shiu RP & Paterson JA 1984 Alterations of cell shape, adhesion,
and lipid accumulation in human breast cancer cells (T-47D)
by human prolactin and growth hormone. Cancer Research 44
1178-1186.
Sicinski P, Donaher JL, Parker SB, Li T, Fazeli A, Gardner H,
Haslam SZ, Bronson RT, Elledge SJ & Weinberg RA 1995
Cyclin D1 provides a link between development and
oncogenesis in the retina and breast. Cell 82 621-630.
Sinha YN 1995 Structural variants of prolactin: occurrence and
physiological significance. Endocrine Reviews 16 354-369.
Sinha YN, DePaolo LV, Haro LS, Singh RNP, Jacobsen BP, Scott
KE & Lewis UJ 1991 Isolation and biochemical properties of
four forms of glycosylated porcine prolactin. Molecular and
Cellular Endocrinology 80 203-213.
Steinmetz RW, Grant AL & Malven PV 1993 Transcription of
prolactin gene in milk secretory cells of the rat mammary
gland. Journal of Endocrinology 36 271-276.
Stevens AN, Wang Y, Sieger KA, Lu H-F & Yu-Lee LY 1995
Biphasic transcription regulation of the interferon regulatory
factor-1 gene by prolactin: involvement of -interferonactivated sequence and stat-related proteins. Molecular
Endocrinology 9 513-525.
Strung I, Gray RA, Rigby HB & Strutton G 1997 Two case
reports of breast carcinoma associated with prolactinoma.
Pathology 29 320-323.
Sutherland RL & Foo MS 1979 Differential binding of
antiestrogen by rat uterine and chick oviduct cytosol.
Biochemical and Biophysical Research Communications 91
183-191.
Sutherland RL, Murphy LC, Foo MS, Green MD, Whybourne
AM & Krozowski ZS 1980 High-affinity anti-oestrogen
binding sites distinct from the oestrogen receptor. Nature 288
273-275.
Sutherland RL, Watts CKW, Hall RE & Ruenitz PC 1987
Mechanism of growth inhibition by nonsteroidal antioestrogens in human breast cancer cells. Journal of Steroid
Biochemistry 27 891-897.

403

Vonderhaar: Prolactin and breast cancer

Tanaka T, Shiu RPC, Gout PW, Beer CT, Noble RL & Friesen HG
1980 A new sensitive and specific bioassay for lactogenic
hormones: measurement of prolactin and growth hormone in
human serum. Journal of Clinical Endocrinology and
Metabolism 51 1058-1063.
Touraine P, Martini JF, Zafrani B, Durand JC, Labaille F, Malet
C, Nicolas A, Trivin C, Postel-Vinay MC, Kuttenn F & Kelly
PA 1998 Increased expression of prolactin receptor gene
assessed by quantitative polymerase chain reaction in human
breast tumors versus normal breast tissue. Journal of Clinical
Endocrinology and Metabolism 83 667-674.
VanderKuur J, Allevato G, Billestrup N, Norstedt G & Carter-Su
C 1995 Growth hormone-promoted tyrosyl phosphorylation
of SHC proteins and SHC association with Grb2. Journal of
Biological Chemistry 270 7587-7593.
Vonderhaar BK 1984 Hormones and growth factors in mammary
gland development. In Control of Cell Growth and
Proliferation, pp 11-33. Ed CM Veneziale. New York: Van
Noostrand, Reinhold and Co., Inc.
Vonderhaar BK 1987 Prolactin: transport, function, and receptors
in mammary gland development and differentiation. In The
Mammary Gland, pp 383-438. Eds MC Neville & CW Daniel.
New York: Plenum Publishing Corporation.
Vonderhaar BK 1998 Prolactin: the forgotten hormone of breast
cancer. Pharmacology and Therapeutics 79 169-178.
Vonderhaar BK 1999 Prolactin and breast cancer: the new
chapter. Women and Cancer 1 (In Press).
Vonderhaar BK & Biswas R 1987 Prolactin effects and regulation
of its receptors in human mammary tumor cells. In Cellular
and Molecular Biology of Mammary Cancer, pp 205-219. Eds
D Medina, W Kidwell, G Hepner & E Anderson. New York:
Plenum Publishing Corp.
Vonderhaar BK & Banerjee R 1991 Is tamoxifen also an
antilactogen? Molecular and Cellular Endocrinology 79
C159-C163.
Wakao H, Gouilleux F & Groner B 1994 Mammary gland factor
(MGF) is a novel member of the cytokine regulated
transcription factor gene family and confers the prolactin
response. EMBO Journal 13 2182-2191.
Walker AM 1994 Phosphorylated and nonphosphorylated
prolactin isoforms. Trends in Endocrinology and Metabolism
5 195-200.
Wang DY, Hampson S, Kwa HG, Moore JW, Bulbrook RD,
Fentiman IS, Hayward JL, King RJB, Millis RR, Rubens RD

404

& Allen DS 1986 Serum prolactin levels in women with breast


cancer and their relationship to survival. European Journal of
Cancer and Clinical Oncology 22 487-492.
Waters MJ, Daniel N, Bignon C & Dijiane J 1995 The rabbit
mammary gland prolactin receptor is tyrosine phosphorylated
in response to prolactin in vivo and in vitro. Journal of
Biological Chemistry 270 5136-5143.
Waters SB & Rillema JA 1989 Role of protein kinase C in the
prolactin-induced responses in mouse mammary gland
explants. Molecular and Cellular Endocrinology 63 159-166.
Watson C & Miller WR 1995 Elevated levels of members of the
STAT family of transcription factors in breast carcinoma
nuclear extracts. British Journal of Cancer 71 840-844.
Welsch CW & Gribler C 1973 Prophylaxis of spontaneously
developing mammary carcinoma in C3H/HeJ female mice by
suppression of prolactin. Cancer Research 33 2939-2946.
Welsch C & Nagasawa H 1977 Prolactin and murine mammary
tumorigenesis: a review. Cancer Research 37 951-963.
Welsch CW, Iturri GC & Brennan MJ 1976 DNA synthesis of
human, mouse, and rat mammary carcinomas in vitro:
influence of insulin and prolactin. Cancer 38 1272-1281.
Wicks JR & Brooks CL 1995 Biological activity of phosphorylated and dephosphorylated bovine prolactin. Molecular
and Cellular Endocrinology 112 223-229.
Wilson GD, Woods KL, Walker RA & Howell A 1980 Effect of
prolactin on lactalbumin production by normal and malignant
breast tissue in organ culture. Cancer Research 40 486-489.
Wong VLY, Compton MM & Witorsch RJ 1986 Proteolytic
modification of rat prolactin by subcellular fractions of the
lactating rat mammary gland. Biochimica et Biophysica Acta
881 167-175.
Yoshiji H, Harris SR & Thorgeirsson UP 1997 Vascular
endothelial growth factor is essential for initial but not
continued in vivo growth of human breast carcinoma cells.
Cancer Research 57 3924-3928.
Young KH, Buhi WC, Horseman N, Davis J, Kraeling R, Linzer
D & Bozer FW 1990 Biological activities of glycosylated and
nonglycosylated porcine prolactin. Molecular and Cellular
Endocrinology 71 155-162.
Zabala MT & Garcia-Ruiz JP 1989 Regulation of expression of
the messenger ribonucleic acid encoding the cytosolic form of
phosphoenolpyruvate carboxykinase in liver and small
intestine of lactating rats. Endocrinology 125 2587-2593.

Вам также может понравиться