02.12.2012 Views

Estrogen Receptor Null Mice - Endocrine Reviews

Estrogen Receptor Null Mice - Endocrine Reviews

Estrogen Receptor Null Mice - Endocrine Reviews

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

0163-769X/99/$03.00/0<br />

<strong>Endocrine</strong> <strong>Reviews</strong> 20(3): 358–417<br />

Copyright © 1999 by The <strong>Endocrine</strong> Society<br />

Printed in U.S.A.<br />

<strong>Estrogen</strong> <strong>Receptor</strong> <strong>Null</strong> <strong>Mice</strong>: What Have We Learned<br />

and Where Will They Lead Us?<br />

JOHN F. COUSE AND KENNETH S. KORACH<br />

<strong>Receptor</strong> Biology Section, Laboratory of Reproductive and Developmental Toxicology, National Institute<br />

of Environmental Health Sciences, National Institutes of Health, Research Triangle Park, North<br />

Carolina 27709<br />

I. Introduction—A Historical Perspective<br />

II. <strong>Estrogen</strong> <strong>Receptor</strong>s (ER)<br />

A. Gene and mRNA structure<br />

B. Mechanism of ER action<br />

C. Generation of the estrogen receptor null mice<br />

III. Reproductive Tract Phenotypes of the Female<br />

A. Uterus<br />

B. Vagina<br />

C. Oviduct<br />

D. Ovary<br />

IV. Mammary Gland<br />

A. �ERKO phenotype<br />

B. �ERKO phenotype<br />

C. ER� and oncogene-induced tumorigenesis: Wnt-<br />

1/�ERKO mice<br />

V. Reproductive Tract Phenotypes of the Male<br />

A. Testicular function and spermatogenesis<br />

B. Accessory sex organs<br />

VI. Neuroendocrine System<br />

A. Hypothalamic-pituitary axis<br />

B. Behavior<br />

VII. Phenotypes in Peripheral Tissues<br />

A. Skeletal system<br />

B. Cardiovascular system<br />

C. Adipogenesis<br />

VIII. Comparison with Human Disease and Models of Deficient<br />

<strong>Estrogen</strong> Action<br />

A. Ovarian carcinogenesis<br />

B. Chronic anovulation<br />

C. ER� and aromatase deficiency<br />

IX. Summary<br />

I. Introduction—A Historical Perspective<br />

NE OF the challenging problems confronting bio-<br />

“O logical scientists has been the manner in which<br />

hormones serve as regulators of biochemical processes in<br />

tissues of higher animals” [E. V. Jensen and E. R DeSombre,<br />

1973 (1).]<br />

At the time of the above statement, the laboratories of<br />

Address reprint requests to: Dr. Kenneth S. Korach, <strong>Receptor</strong> Biology<br />

Section, Laboratory of Reproductive and Developmental Toxicology,<br />

National Institute of Environmental Health Sciences, National Institutes<br />

of Health, MD B3–02, P.O. Box 12233, Research Triangle Park, North<br />

Carolina 27709 USA.<br />

358<br />

Elwood Jensen and Jack Gorski had spent 10 yr providing<br />

experimental evidence to support the concept of an intracellular<br />

“receptor” protein for steroid hormones. Their combined<br />

work had even led to a proposed model by which the<br />

interactions of the receptor and the steroid were involved in<br />

mediating the cellular effects of the hormone (1). The first of<br />

these receptors to be characterized was for the female sex<br />

hormone, 17�-estradiol (E 2) (2, 3). Since this time, similar<br />

receptors for testosterone, progesterone, glucocorticoids,<br />

thyroid hormone, vitamin-D 3, and retinoids have been discovered<br />

and now form a portion of a large family of nuclear<br />

hormone receptors (4). Although significant strides have<br />

been made since, Jensen’s introductory statement in a review<br />

published more than 25 yr ago is still contemporary with the<br />

current research goals toward understanding the growing<br />

family of nuclear receptors. This is not to say that little<br />

progress has been made, which is quite to the contrary, but<br />

rather to state that although advances in technology and<br />

molecular biology have allowed for extensive insight, there<br />

is still a great deal to be learned as we move into the next<br />

century.<br />

Our current understanding of the various roles of the<br />

steroid, thyroid, and retinoid hormones in development and<br />

normal physiology and the mechanisms by which these actions<br />

are mediated is due, in large part, to the generation of<br />

a series of reagents and tools over the past 40 yr. The first of<br />

these was the synthesis and use of tritium-labeled E 2 with<br />

high specific activity, allowing for the first reports of the<br />

detection and simple characterization of an estrogen-binding<br />

component, or “estrophilin” (1, 3, 5). This protein exhibited<br />

a binding affinity for estradiol that was several fold higher<br />

compared with the other gonadal steroids and was found<br />

only in tissues previously shown to respond to estradiol in<br />

terms of growth and increased RNA synthesis (1, 3). These<br />

studies described the uptake and concentration of radiolabeled<br />

estradiol by a specific protein unique to the cells of the<br />

uterus, vagina, and pituitary and thereby disputed the current<br />

thought that estrogen action required enzymatic metabolism<br />

of the hormone (2, 3, 6). Thus, the concept of a<br />

hormone receptor protein in target tissues was initiated. Autoradiographic<br />

studies with radiolabeled ligand further<br />

demonstrated the strong association of estradiol with the<br />

nuclei of cells lining the rat uterus within 4 h after injection<br />

(7). These studies were significantly advanced by pharmacological<br />

approaches with early nonsteroidal estrogen antagonists,<br />

the first being the triphenylethylene, MER-25 (2),


June, 1999 ESTROGEN RECEPTOR NULL MICE 359<br />

followed soon after by a series of similar compounds, e.g.,<br />

clomiphene, nafoxidine, CI-628, and tamoxifen (1, 8). These<br />

reagents were previously known to inhibit the uterotropic<br />

effects of estradiol in a dose-dependent manner but, more<br />

importantly, were later shown to block estradiol uptake and<br />

binding in target tissues. Still, at this time the exact role that<br />

the hormone-receptor complex played in the final manifestation<br />

of the hormonal effects remained unclear. It was suggested<br />

that the receptor may simply fulfill a “transport” role<br />

to move the steroid hormone from the cytoplasm to the<br />

nucleus of the cell (1). Nonetheless, from these early studies,<br />

the definition of an estrogen target tissue now included not<br />

only the exhibition of a measurable response to the natural<br />

hormone (E 2) but also one that possessed detectable levels of<br />

the estrogen receptor (ER).<br />

Much of the early characterization of the ER relied on the use<br />

of sucrose gradient analysis of nuclear and cytoplasmic cell<br />

fractions under varied salt concentrations (3). This procedure,<br />

along with gel electrophoresis and filtration using radiolabeled<br />

estradiol, allowed for the generation of semipure fractions of the<br />

estradiol-binding protein. These preparations led to the generation<br />

of antisera specific to the ER protein, another notable step<br />

in receptor research (9). The later development of monoclonal<br />

antibodies to the ER and the immunohistochemical techniques<br />

that soon followed provided evidence of the predominantly<br />

nuclear localization of the receptor protein in target cells (10, 11).<br />

The late 1970s brought numerous reports of the tissue distribution<br />

and localization of the ER in humans and laboratory<br />

animals, confirming much of the findings of earlier steroid<br />

autoradiography studies (7, 11). The antibodies were also used<br />

to further purify large amounts of the ER from target tissues,<br />

allowing for more detailed studies of the receptor structure and<br />

function.<br />

The 1980s witnessed the cloning and sequencing of the<br />

cDNAs for several of the steroid hormone receptors, which<br />

proved to be a seminal step toward understanding their<br />

mechanisms of action. The first cDNA to be cloned for a<br />

member of the nuclear receptor family was that for the human<br />

glucocorticoid receptor (12), followed soon after with<br />

the description of the human ER cDNA (13, 14). Since this<br />

time, the cDNAs encoding several other members of the<br />

nuclear receptor family have been described (4). The current<br />

list of isolated ER cDNAs includes those for the chicken (15),<br />

mouse (16), rat (17), Xenopus laevis (18), and rainbow trout<br />

(19). Sequence analysis of the various receptor cDNAs demonstrated<br />

a high degree of similarity and led to their inclusion<br />

in a superfamily of nuclear receptors possessing a defined<br />

motif of functional domains (4).<br />

A vital contribution from the cloning of the nuclear receptor<br />

cDNAs was the ability to express the receptor proteins<br />

in in vitro mammalian (20) or yeast (21) cell systems. These<br />

techniques, as well as cell-free in vitro transcription (22),<br />

allowed detailed characterizations of the different functional<br />

domains of the receptor. Recombinant DNA methodologies<br />

provided for the construction of receptor cDNAs possessing<br />

precise truncations, deletions, point mutations, or additional<br />

sequences. These powerful techniques led to the in vitro<br />

expression of chimeric and mutant receptors and great advances<br />

in the dissection and mapping of the specific domains<br />

and distinct residues critical to receptor function (21–25). The<br />

late 1980s witnessed multiple descriptions of naturally occurring<br />

variants and mutations of several of the nuclear<br />

receptor transcripts (26), including the ER (27–29). Their<br />

identification and functional characterization have led to<br />

further insight into the mechanisms of action of specific<br />

domains of the receptor proteins as well as the receptors as<br />

a whole. Furthermore, the detection of these nuclear receptor<br />

transcript variants in vivo has allowed speculation concerning<br />

their possible roles in alternate or abnormal hormonal<br />

signaling in normal and neoplastic tissues (27).<br />

By the 1990s, it was evident that the members of the steroid/<br />

thyroid hormone superfamily of receptors were intracellular<br />

proteins that functioned in the nucleus to regulate transcription<br />

of target genes. This growing family of receptors now includes<br />

those for the sex and adrenal steroids, thyroid hormones, retinoids,<br />

vitamin D 3, and eicosinoids (reviewed in Refs. 4 and 30).<br />

The inclusion of nuclear receptor-like proteins with no known<br />

ligand, termed orphan receptors, such as those for the chicken<br />

ovalbumin upstream promoter (COUP), and steroidogenic factor-1<br />

(SF-1), has expanded the family to now include approximately<br />

150 distinct proteins (30).<br />

Our knowledge of the expression patterns and mechanism<br />

of action of the nuclear receptors has led to a greater appreciation<br />

of their involvement in normal physiology and disease.<br />

The current decade has witnessed several advances in<br />

our understanding of the molecular biology and overall<br />

physiological role of these proteins. Especially significant to<br />

these efforts has been the discovery and/or development of<br />

three distinct aspects. The first of these was the discovery of<br />

coregulators, a second group of nuclear proteins that further<br />

modulate the actions of unoccupied as well as ligand (agonist<br />

or antagonist)-bound receptors (reviewed in Ref. 31). The<br />

continued characterization of the coregulator proteins has<br />

provided some insight toward the long sought explanations<br />

for the cell- and tissue-specific mixed agonist/antagonist<br />

activity of certain ligands (31). The second of these advances<br />

was the generation of crystal structures for the ligand-binding<br />

domains of the retinoic acid receptor-� (32), retinoic X<br />

receptor-� (33), thyroid receptor-� (34), and ER� (35, 36).<br />

These new data continue to allow for the description of the<br />

long speculated conformational changes that result in the<br />

receptor after binding of the natural ligand, and how these<br />

changes may differ from those induced by natural or synthetic<br />

agonists and antagonists.<br />

The third development of great impact in this decade has<br />

been the use of gene-targeting technology and transgenic techniques<br />

to disrupt the genes encoding several members of the<br />

steroid/thyroid hormone receptor superfamily. This methodology<br />

has allowed for the generation of transgenic mice that lack<br />

a functional gene for a specific receptor, as well as germline<br />

passage of this mutation. Hormone resistance due to naturally<br />

occurring mutations in the genes encoding the receptors for<br />

androgen (37, 38), glucocorticoids (39), and thyroid hormones<br />

(40) had already been described in humans and laboratory<br />

animals and thereby provided great insight into the role of these<br />

hormones in development and normal physiology. However,<br />

similar mutations had not yet been reported for other members<br />

of the superfamily of nuclear receptors. The impact of the genetargeting<br />

technique is evident in the generation of several<br />

“knockout” mice for the nuclear receptors, including mice lack-


360 COUSE AND KORACH Vol. 20, No. 3<br />

ing multiple forms of the retinoic acid receptors (41–43), the<br />

progesterone receptor (44), the vitamin D 3 receptor (45), the ERs<br />

(46, 47), and the coregulator protein, steroid receptor coactivator-1<br />

(48). If nothing else, the naturally occurring mutants combined<br />

with the generation of the knockout models, have<br />

confirmed the basic hypotheses put forth almost 30 yr ago,<br />

i.e., that the receptor proteins found to tightly and specifically<br />

bind a hormone within a tissue are indeed critical<br />

for mediating the biological effect of the hormone within<br />

the target cell (1).<br />

This review will discuss one such advance originating<br />

from the gene-targeting boom that occurred in the 1990s, the<br />

ER knockout mice. The broad content of this review is a<br />

reflection of the current state of research in the expanding<br />

field of estrogen action, as illustrated by the discovery of a<br />

second ER, the ER�. A large extent of the discussion will<br />

focus on the mice lacking the classical ER, the ER� knockout<br />

mice (�ERKO). The reasons for this are 2-fold: 1) the �ERKO<br />

mouse has been available for more than 6 yr, compared with<br />

less than 1 yr for the �ERKO, and therefore has been more<br />

thoroughly studied and characterized; 2) the phenotypes of<br />

the �ERKO mice appear to be more broad in nature compared<br />

with those of the less well studied �ERKO mice, although<br />

continued research may lessen this disparity. The<br />

extent to which the ERKO models have been used in various<br />

fields of biological research is illustrated in Table 1. This<br />

review will discuss the phenotypes of the ERKO mice and the<br />

multiple ways in which these models continue to influence<br />

the field of estrogen research. Foremost, several of the hypotheses<br />

of estrogen and ER action put forth from the efforts<br />

of numerous investigators over the years were confirmed by<br />

our observations in the ERKO. Furthermore, certain phenotypes<br />

have introduced unforeseen critical roles of estrogen in<br />

some physiological systems, such as in male fertility. Later<br />

years have witnessed the use of the �ERKO as a research tool<br />

to investigate specific biochemical pathways and neoplasia<br />

in the absence of estrogen action. Where applicable, we will<br />

contrast the phenotypes of the two ERKO models, as well as<br />

compare the relevant phenotypes of knockout models for<br />

other hormone signaling systems. And finally, we will discuss<br />

how these models compare with the few cases of insufficient<br />

estrogen synthesis and the single reported case of<br />

estrogen insensitivity in humans. Before we continue, however,<br />

we wish to take this opportunity to establish consistency<br />

in the abbreviations used to refer to the ER-knockout<br />

(ERKO) mice in the literature. Although the terms ER�KO<br />

and ER�KO have appeared in previous reports from our own<br />

as well as other laboratories, we propose that the abbreviations<br />

used above and throughout the remainder of this review,<br />

i.e., �ERKO and �ERKO, be the consensus abbreviations<br />

used hereafter to refer to the two models.<br />

A. Gene and mRNA structure<br />

II. <strong>Estrogen</strong> <strong>Receptor</strong>s<br />

For several years it was thought that only a single form of the<br />

nuclear ER existed. However, in 1996, multiple laboratories<br />

independently reported the discovery of a second type of ER in<br />

the rat (49), mouse (50), and human (51). This newly discovered<br />

receptor was termed ER�, resulting in the classical ER being<br />

referred to as ER�. The two receptors are not isoforms of each<br />

other, but rather distinct proteins encoded by separate genes<br />

located on different chromosomes. As a result of this discovery,<br />

we now know that the ER shares a phenomenon of multiple<br />

forms previously described for other members of the nuclear<br />

hormone superfamily, including the receptors for thyroid hormone,<br />

retinoids, mineralocorticoids, and progesterone (26, 30).<br />

A detailed description of the structure and mechanism of action<br />

of the ERs is beyond the scope of this review, and therefore only<br />

a brief discussion of the relevant points will appear here. For<br />

more detailed descriptions of the mechanism of steroid receptor-regulated<br />

gene transcription, readers are encouraged to<br />

seek several recent reviews (52–55).<br />

Transcription of the mouse ER� gene in vivo predominantly<br />

results in a single transcript of approximately 6.3 kb transcribed<br />

from 9 exons. This transcript encodes a protein of 599 amino<br />

acids with an approximate molecular mass of 66 kDa (16). The<br />

human ER� is slightly shorter at 595 amino acids but exhibits<br />

a similar molecular mass (13, 14). Whereas the human ER� gene<br />

has been mapped to chromosome 6 (56), the mouse ER� gene<br />

is located on chromosome 10 (57). The existence of multiple<br />

promoter and regulatory regions in the 5�-untranslated sequences<br />

of the human and rat ER� have been described, but<br />

only a single open reading frame appears to exist (58–60).<br />

Numerous reports have described the discovery and characterization<br />

of naturally occurring variants and mutations of the<br />

ER� mRNA in normal as well as neoplastic tissues of several<br />

species (reviewed in Refs. 27, 28, and 61). Although the existence<br />

of true protein products of these ER� mRNA variants in<br />

vivo remains controversial, their transactivational activities<br />

in in vitro cell culture systems have been intensely<br />

described and have furthered our understanding of the<br />

functional domains of the receptor (27, 61).<br />

Initial studies indicated that the rodent ER� was composed<br />

of 485 amino acids and an estimated molecular mass<br />

of 54 kDa and therefore was slightly smaller than the ER� (49,<br />

50, 62). The majority of this difference in size between the two<br />

ERs was due to a significantly shorter N� terminus in the<br />

deduced ER� protein. Unlike the ER� gene, Northern blot<br />

analyses of ovarian RNA from both the rat and mouse indicates<br />

the presence of multiple ER� transcripts (50, 63, 64).<br />

Furthermore, open reading frames initiating up-stream from<br />

those originally described have now been discovered in the<br />

mouse, rat, bovine, and human ER� mRNA (65–69). These<br />

studies have also provided evidence to support that the<br />

upstream start codons are the likely initiation sites of<br />

translation and therefore suggest the possibility of an ER�<br />

protein of 527–530 amino acids and a calculated molecular<br />

mass of approximately 60 kDa (65–67, 69). However, convincing<br />

Western blots from tissue extracts to indicate the<br />

true in vivo molecular mass of the ER� have remained<br />

difficult to produce with the antibody preparations currently<br />

available. Similar to ER�, a number of variants of the<br />

ER� mRNA have already been described. These include a<br />

conserved insertion of 18 amino acids in a C�-terminal<br />

region of the ER� in the rat (70, 71), human, and mouse<br />

(72), the deletion of one or more exons in these same<br />

species (70, 72, 73), and various isoforms in the extreme<br />

C�-terminus of the human ER� (68).


June, 1999 ESTROGEN RECEPTOR NULL MICE 361<br />

TABLE 1. Published reports involving the �ERKO and �ERKO models and the ER�-deficient human<br />

Author<br />

ER-� knockout mice: general<br />

Yr. Title Ref.<br />

Lubahn et al. 1993 Alteration of reproductive function but not prenatal sexual development after insertional<br />

disruption of the mouse estrogen receptor gene.<br />

46<br />

Couse et al. 1995 Analysis of transcription and estrogen insensitivity in the female mouse after targeted<br />

disruption of the estrogen receptor gene.<br />

123<br />

Couse et al.<br />

ER-� knockout mice: general<br />

1997 Tissue distribution and quantitative analysis of estrogen receptor-� (ER�) and estrogen-�<br />

(ER�) messenger ribonucleic acid in the wild-type and ER�-knockout mouse.<br />

93<br />

Krege et al.<br />

Female reproductive tract<br />

1998 Generation and reproductive phenotypes of mice lacking estrogen receptor-�. 47<br />

Curtis et al. 1996 Physiological coupling of growth factor and steroid receptor-signaling pathways: estrogen<br />

receptor knockout mice lack estrogen-like response to epidermal growth factor.<br />

170<br />

Lindzey et al. 1996 Uterotropic effects of dihydrotestosterone in estrogen receptor knockout and wild-type mice. a<br />

154<br />

Das et al. 1997 <strong>Estrogen</strong>ic responses in estrogen receptor-� deficient mice reveal a distinct signaling pathway. 173<br />

Cooke et al. 1997 Stromal estrogen receptors mediate mitogenic effects of estradiol on uterine epithelium. 171<br />

Buchanan et al. 1998 Role of stromal and epithelial estrogen receptors in vaginal epithelial proliferation,<br />

stratification, and cornification.<br />

198<br />

Ghosh et al. 1998 Methoxychlor acts in ER�-KO mice through an ER-� independent mechanism. a<br />

174<br />

Rosenfeld et al. 1998 <strong>Estrogen</strong> receptor-� knockout mice reveal a role for estrogen in mammalian female sexual<br />

development. a<br />

494<br />

Kurita et al. 1998 <strong>Estrogen</strong> induces progesterone receptors in uterine stroma of estrogen receptor � knockout<br />

mouse. a<br />

495<br />

Buchanan et al. 1999 Tissue compartment-specific estrogen receptor-� participation in the mouse uterine epithelial<br />

secretory response.<br />

172<br />

Curtis et al. 1999 Disruption of estrogen signaling does not prevent progesterone action in the estrogen receptor<br />

knockout mouse uterus.<br />

183<br />

Schomberg et al.<br />

Mammary gland<br />

1999 Targeted disruption of the estrogen receptor-� (ER�) gene in mice: characterization of ovarian<br />

responses and phenotypes.<br />

142<br />

Bocchinfuso and Korach 1997 Mammary gland development and tumorigenesis in estrogen receptor knockout mice. 269<br />

Cunha et al. 1997 Elucidation of a role of stromal steroid hormone receptors in mammary gland growth and<br />

development by tissue recombination experiments.<br />

278<br />

Kenney et al. 1997 The response of null estrogen receptor mammary parenchyma and mesenchyme in vivo. 496<br />

Bocchinfuso et al.<br />

Male reproductive tract<br />

1999 An MMTV-Wnt-1 transgene induces mammary gland hyperplasia and tumorigenesis in mice<br />

lacking estrogen receptor-�.<br />

297<br />

Eddy et al. 1996 Targeted disruption of the estrogen receptor gene in male mice causes alteration of<br />

spermatogenesis and infertility.<br />

315<br />

Donaldson et al. 1996 Morphometric study of the gubernaculum in male estrogen receptor mutant mice. 316<br />

Hess et al. 1997 A role for oestrogens in the male reproductive system. 327<br />

Rosenfeld et al.<br />

Neuroendocrine system<br />

1998 Transcription and translation of estrogen receptor-beta in the male reproductive tract of<br />

estrogen knockout and wild-type mice.<br />

121<br />

Scully et al. 1997 Role of estrogen receptor � in the anterior pituitary gland. 282<br />

Shughrue et al. 1997 Responses in the brain of estrogen receptor �-disrupted mice. 400<br />

Shughrue et al. 1997 The disruption of estrogen receptor-� mRNA in forebrain regions of the estrogen receptor-�<br />

knockout mouse.<br />

352<br />

Simerly et al. 1997 <strong>Estrogen</strong> receptor-dependent sexual differentiation of dopaminergic neurons in the preoptic<br />

region of the mouse.<br />

405<br />

Young et al. 1998 <strong>Estrogen</strong> receptor � is essential in the induction of oxytocin receptor by estrogen. 497<br />

De Vries et al. 1997 Steroid-responsive vasopressin innervation in the brain of estrogen receptor-�-minus mice. a<br />

498<br />

Lindzey et al. 1998 Effects of castration and chronic steroid treatments on hypothalamic gonadotropin-releasing<br />

hormone content and pituitary gonadotropins in male wild-type and estrogen receptor-�<br />

knockout mice.<br />

317<br />

Lindzey et al. 1998 Steroid regulation of gonadotrope function in female wild-type (WT) and estrogen receptor-�<br />

knockout (ERKO) mice. a<br />

252<br />

Ogawa et al. 1998 Forebrain steroid receptor immunoreactive cells in neonatal and prepubertal estrogen receptor-<br />

� gene deficient (ERKO) mice. a<br />

426<br />

Moffatt et al. 1998 Induction of progestin receptors by estradiol in the forebrain of estrogen receptor-� genedisrupted<br />

mice.<br />

423<br />

Wersinger et al. 1998 Estradiol decreases proGnRH immunoreactivity in female wild type but not estrogen receptor �<br />

knockout mice. a<br />

499<br />

B. Mechanism of ER action<br />

The ERs are classified as class I members of the superfamily<br />

of nuclear hormone receptors, defined as a ligandinducible<br />

transcription factor (30). Early studies indicated the<br />

ER was cytoplasmic and became localized to the nucleus only<br />

upon ligand binding, providing the basis of the initial “twostep<br />

mechanism” of hormone action (1, 3). However, it is now<br />

accepted that the ER is a predominantly nuclear protein<br />

regardless of whether or not it is complexed with ligand (74).<br />

The inactive ER exists in a complex consisting of several


362 COUSE AND KORACH Vol. 20, No. 3<br />

TABLE 1. continued<br />

Author Yr. Title Ref.<br />

Gu et al.<br />

Behavior<br />

1999 Rapid action of 17�-estradiol on kainate-induced currents in hippocampal neurons lacking<br />

intracellular estrogen receptors.<br />

364<br />

Ogawa et al. 1996 Reversal of sex roles in genetic female mice by disruption of estrogen receptor gene. 417<br />

Ogawa et al. 1997 Behavioral effects of estrogen receptor gene disruption in male mice. 318<br />

Rissman et al. 1997 <strong>Estrogen</strong> receptors are essential for female sexual receptivity. 419<br />

Wersinger et al. 1997 Masculine sexual behavior is disrupted in male and female mice lacking a functional estrogen<br />

receptor � gene.<br />

427<br />

Fugger et al. 1998 <strong>Estrogen</strong> receptor � is not required for estradiol’s effects on inhibitory avoidance behavior in<br />

female mice. a<br />

500<br />

Fugger et al. 1998 Characterization of spatial behavior and hippocampal electrophysiology in male estrogen<br />

receptor-�-minus mice. a<br />

501<br />

Fugger et al. 1998 Sex differences in the activational effect of ER� on spatial learning. 502<br />

Ogawa et al. 1998 Roles of estrogen receptor-� gene expression in reproduction-related behaviors in female mice. 418<br />

Ogawa et al.<br />

Cardiovascular system<br />

1998 Modifications of testosterone-dependent behaviors by estrogen receptor-� gene disruption in<br />

male mice.<br />

428<br />

Johns et al. 1996 Disruption of estrogen receptor gene prevents 17�-estradiol-induced angiogenesis in transgenic<br />

mice.<br />

503<br />

Iafrati et al. 1997 <strong>Estrogen</strong> inhibits the vascular injury response in estrogen receptor-� deficient mice. 464<br />

Johnson et al. 1997 Increased expression of the cardiac L-type channel in estrogen receptor-deficient mice. 472<br />

Rubanyi et al. 1997 Vascular estrogen receptors and endothelium-derived nitric oxide production in the mouse<br />

aorta: gender difference and effect of estrogen receptor gene disruption.<br />

467<br />

Srivastava et al. 1997 <strong>Estrogen</strong> up-regulates apolipoprotein E (Apo E) gene expression by increasing Apo E mRNA in<br />

the translating pool via the estrogen receptor �-mediated pathway.<br />

460<br />

Freay et al. 1997 Mechanism of vascular smooth muscle relaxation by estrogen in depolarized rat and mouse<br />

aorta.<br />

504<br />

Cross et al. 1998 <strong>Estrogen</strong> receptor knock-out mice exhibit higher myocardial contractility than wild-type mice<br />

but are equally susceptible to ischemic injury. a<br />

Bone<br />

505<br />

Roemmich et al. 1997 Bone strain independently augments bone mineral in pubertal mice with the disrupted<br />

estrogen receptor gene. a<br />

506<br />

Pan et al. 1997 <strong>Estrogen</strong> receptor-alpha knockout (ERKO) mice lose trabecular and cortical bone following<br />

ovariectomy. a<br />

448<br />

Korach et al.<br />

Other<br />

1997 The effects of estrogen receptor gene disruption on bone. 447<br />

Smithson et al. 1997 The role of estrogen receptors and androgen receptors in sex steroid regulation of B<br />

lymphopoiesis.<br />

507<br />

Hurn et al. 1998 Stroke in mice deficient in classical estrogen receptors (ERKO�). a<br />

508<br />

Couse et al. 1998 Use of the estrogen receptor-� to investigate the role of ER� in the developmental and<br />

carcinogenic actions of diethylstilbestrol. a<br />

509<br />

Taylor and Lubahn 1998 Impaired glucose tolerance in the ER�KO mouse. a<br />

490<br />

Yellayi et al. 1998 Role of estrogen receptor-� (ER�) in the development and function of the thymus in male and<br />

female mice. a<br />

ER-deficient humans<br />

510<br />

Smith et al. 1994 <strong>Estrogen</strong> resistance caused by a mutation in the estrogen receptor gene in a man. 116<br />

Sudhir et al. 1997 Endothelial dysfunction in a man with disruptive mutation of the oestrogen-receptor gene. 492<br />

Sudhir et al. 1997 Premature coronary artery disease associated with a disruptive mutation in the estrogen<br />

receptor gene in a man.<br />

493<br />

Dieudonne et al. 1998 Immortalization and characterization of bone marrow stromal fibroblasts from a patient with a<br />

loss of function mutation in the estrogen receptor-� gene.<br />

511<br />

a Abstract.<br />

heat-shock and other proteins that appear to disassociate<br />

upon ligand binding, resulting in a “transformation” of the<br />

receptor to an active state (75). With continued research, the<br />

two-step mechanism model has evolved to state that upon<br />

binding of estradiol, or an estrogenic ligand, the transformed<br />

receptors form dimers that tightly associate with specific<br />

consensus DNA sequences, consisting of 15-bp inverted palindromes<br />

in the regulatory regions of target genes (52, 74–<br />

76). This complex then interacts with basal transcription factors,<br />

coregulator proteins, and other transcription factors to<br />

ultimately regulate transcription of the target gene (52, 55,<br />

76). However, in recent years, pathways of gene activation by<br />

the steroid receptors that deviate from this classical model<br />

have been described. These include gene activation by<br />

ligand-bound steroid receptors without evidence of direct<br />

DNA binding, but rather via interaction with other DNAbound<br />

transcription factors, such as an AP-1 complex (77, 78).<br />

In addition, ligand-independent activation of the receptor<br />

through pathways that alter the activity of cellular kinases<br />

and phosphatases has been demonstrated both in vitro and<br />

in vivo (reviewed in Ref. 55). The discovery of these pathways<br />

strongly supports the great importance of the ER and its<br />

ability to possibly provide diverse physiological functions<br />

even in the absence of ligand.


June, 1999 ESTROGEN RECEPTOR NULL MICE 363<br />

The ER� and ER� proteins are composed of six functional<br />

domains, labeled A–F, a signature characteristic of members<br />

of the superfamily of steroid/thyroid hormone nuclear receptors<br />

(Fig. 1). The N�-terminal A/B domain is the least<br />

conserved among all members and demonstrates only 17%<br />

identity between the human ER� and ER� (64). In contrast,<br />

the C domain is the most highly conserved among the different<br />

members of the family. It possesses two zinc fingers<br />

forming a helix-loop-helix motif and primarily functions in<br />

tightly binding the receptor to the DNA hormone response<br />

elements. The sequences encoding the two zinc fingers possess<br />

97% homology between the ER� and ER� genes and are<br />

located in separate exons (exons 3 and 4) in each (50, 64, 79).<br />

The E domain, or ligand-binding domain, confers ligand<br />

specificity to the receptor and is moderately conserved<br />

among the members of the superfamily. The ER� and ER�<br />

proteins possess 60% conservation of the residues in the E<br />

domain; however, each binds estradiol with nearly equal<br />

affinity and exhibits a very similar binding profile for a large<br />

number of natural and synthetic ligands (80). The D domain<br />

possesses signals for nuclear localization of the receptor and<br />

exhibits approximately 30% identity between the two human<br />

forms of ER (64). The C�-terminal F domain is unique to the<br />

ER among the nuclear receptors for the gonadal and adrenal<br />

hormones (6) but is not well conserved among the ERs of<br />

different species nor between the ER� and ER�, which share<br />

approximately 18% homology (64). Studies using forms of<br />

the ER� missing the C� terminus have indicated a role for the<br />

F domain in modulating transactivational activity of the ER�<br />

when complexed with mixed agonist/antagonist ligands,<br />

possibly via influencing coregulatory function and/or<br />

dimerization of the receptor (81, 82).<br />

There are also functional domains that span those boundaries<br />

described above. Residues involved in the dimerization<br />

of the receptors are located in the second zinc finger of the<br />

C domain as well as in the major dimerization surface in the<br />

E domain (83, 84). Furthermore, two domains critical to the<br />

transactivational function of the ER are the AF-1 in the N�<br />

terminus and AF-2 in the C� terminus. These two domains<br />

may function independently or interact during the process of<br />

transactivation, depending on the cell type, target promoter,<br />

and the presence and/or type of ligand (52). The AF-2 domain<br />

is critical to the ligand-dependent transactivational activity<br />

of the receptor and may be involved in the recruitment<br />

of coregulator proteins, whereas the AF-1 is thought to be a<br />

region of site-specific phosphorylation involved in ligandindependent<br />

activity of the receptor (reviewed in Refs. 31<br />

and 52). Recent studies have also suggested the presence of<br />

a third domain, AF-2a, within the ligand-binding domain of<br />

the human ER� (85).<br />

The discovery of the ER� has introduced a new level of<br />

complexity to the current model as well. To date, there exist<br />

no data indicating a physiological response solely mediated<br />

by ER�. In contrast, the �ERKO mouse has confirmed the<br />

requirement for ER� in mediating several actions of estradiol,<br />

as will be discussed in this review. Nevertheless, in vitro<br />

experiments from several laboratories have indicated the<br />

possibility of cooperative activity between the two receptors,<br />

FIG. 1. Drawing of the mouse ER proteins, cDNAs, and genes as well as the targeting scheme employed to generate the ERKO mice via<br />

homologous recombination. Shown are the common functional domains of the ER� and ER� receptor proteins, indicating the residues involved<br />

in DNA and ligand binding. The common structure of the cDNAs and genomic genes for the ERs is illustrated, indicating the exon sequences<br />

that encode the functional domains of the receptor. Generation of the �ERKO mouse involved the targeted insertion of a 1.8-kb NEO sequence<br />

into exon 2 of the ER� gene such that the translational reading frame (indicated by the direction of the arrow) of the genes was the same (see<br />

Ref. 46). Generation of the �ERKO mouse involved a similar scheme, in which a 1.8-kb NEO sequence was inserted into exon 3 of the ER� gene;<br />

however, in this case the NEO gene is in the reverse orientation (see Ref. 47). The schematic drawing of the genomic DNA was adapted from<br />

that of the human ER� gene (Ref. 79). Drawing is not to scale.


364 COUSE AND KORACH Vol. 20, No. 3<br />

acting in the form of heterodimers (50, 62, 65, 86). These<br />

studies generally report a tendency of ER� to form homodimers<br />

whereas ER� prefers to heterodimerize with ER�.<br />

However, Giguere et al. report (87) that the heterodimer is the<br />

preferred state when both mouse ERs are present. The transactivational<br />

activity of the heterodimer when assayed in in<br />

vitro mammalian cell transfection assays appears to lie between<br />

that of the more active ER� homodimer and the less<br />

active ER� homodimer (50, 62, 86). A major consideration<br />

when evaluating the possible physiological functions of an<br />

ER�/ER� heterodimer is evidence of coexpression of the two<br />

receptors in the same cell, which has not yet been definitively<br />

reported (88). To this end, studies and reagents are only now<br />

becoming available to directly assess this question.<br />

Several functional characteristics of the two ERs are similar.<br />

The residues critical to function of the AF-2 domain<br />

appear to be identical in the mouse ER� and ER� (87). Tremblay<br />

et al. (50, 89) demonstrated that a tyrosine residue critical<br />

to the function of the AF-2 domain was conserved in both the<br />

ER� and ER� and that mutation of this amino acid resulted<br />

in similar constitutive, ligand-independent transactivational<br />

activity in both receptors. In contrast, the N�-terminal AF-1<br />

domain shows no significant regions of similarity between<br />

the two ERs (87). However, a potential activation site of the<br />

mitogen-activated protein (MAP)-kinase pathway previously<br />

shown for ER� is present and active in the ER� (50, 89).<br />

Additionally, when acting on a basal promoter linked to a<br />

consensus estrogen response element, both ER� and ER�<br />

were able to recruit the coactivator SRC-1 and were equally<br />

susceptible to inhibition by the antiestrogens raloxifene, ICI<br />

164,384, and EM-800 (50).<br />

However, as studies continue, distinct differences at the<br />

molecular level and in the transactivational capacities between<br />

ER� and ER� have been described. Two separate<br />

studies have demonstrated the specificity of the agonist activity<br />

of 4-hydroxytamoxifen to be unique to ER�, although<br />

this appears to be highly dependent on the cell and promoter<br />

context as well as experimental design (50, 90). Furthermore,<br />

Paech et al. (91) reported that when interacting with DNAbound<br />

AP-1 transcription factors, the in vitro transactivational<br />

activity of estrogen agonists and antagonists was quite<br />

different depending on which form of ER was present.<br />

Whereas antagonists, such as raloxifene, tamoxifen, and ICI<br />

164,384, were able to block the stimulatory activity of the<br />

ER�/AP-1 complex, these same compounds acted as potent<br />

agonists when bound to an ER�/AP-1 complex (91). Further<br />

experimental support for the existence of distinct structural<br />

and functional differences between ER� and ER� was recently<br />

provided by Sun et al., who showed that certain nonsteroidal<br />

ligands were receptor selective in their binding and<br />

agonist/antagonist activities (92).<br />

Perhaps the most significant disparity lies in the tissue<br />

distribution of the two receptors. Studies employing the techniques<br />

of RT-PCR and/or ribonuclease protection assay<br />

(RPA) have indicated that ER� mRNA is predominant in the<br />

uterus, mammary gland, testis, pituitary, liver, kidney, heart,<br />

and skeletal muscle, whereas ER� transcripts are significantly<br />

expressed in the ovary and prostate (Fig. 2) (63, 80, 93,<br />

94). These same studies have indicated relatively equal levels<br />

of mRNA for the two receptors in the epididymis, thyroid,<br />

adrenals, bone, and various regions of the brain (80, 93,<br />

95–97). However, as more studies are reported, several discrepancies<br />

in the expression patterns of ER� and ER� among<br />

different species are becoming apparent (80, 93, 96). For<br />

example, whereas ER� mRNA is easily detectable in the<br />

pituitary of the rat (70, 98, 99), human (100), and rhesus<br />

monkey (96), levels in the pituitary of the mouse appear low<br />

to undetectable (93). A similar difference in expression is<br />

apparent in the mammary gland, in which normal and neoplastic<br />

human tissue and cell lines express detectable ER�<br />

mRNA (64, 68, 73, 101, 102), although the mammary gland<br />

of the mouse appears to predominantly express ER� (93).<br />

Furthermore, even in those tissues expressing both ERs, there<br />

is often a distinct expression pattern within the heterogeneous<br />

cell types composing the tissue. In the ovary, ER� is<br />

apparently localized to the granulosa cells of maturing follicles,<br />

whereas ER� is detectable in the surrounding thecal<br />

cells (69, 103, 104). In the prostate of the rat, expression of ER�<br />

and ER� is detectable in the stroma and epithelium, respectively,<br />

but does not appear to be colocalized in any portion<br />

of the tissue (49). However, through the combined use of<br />

immunohistochemistry and in situ hybridization, Shughrue<br />

FIG. 2. RT-PCR for ER� and ER� mRNA in various tissues of the wild-type mouse. RT-PCR was carried out on 0.5 �g of total RNA pooled from<br />

adult wild-type mice using primers specific for the mouse ER� and ER� transcripts (see Refs. 93 and 123). Equal amounts of the individual<br />

RT-PCR reactions were then fractionated on an agarose gel. Note the broad tissue distribution of ER� mRNA, whereas ER� transcripts are<br />

primarily expressed in the ovary, hypothalamus, lung, and male reproductive tract. RT-PCR for �-actin was carried out as a positive control.<br />

(�RT) indicates a negative control, i.e., PCR on total ovarian RNA minus reverse transcriptase, indicating the specificity of the ER primers for cDNAs<br />

generated by the reverse transcriptase enzyme.


June, 1999 ESTROGEN RECEPTOR NULL MICE 365<br />

et al. (88) have demonstrated colocalization of ER� and ER�<br />

to select regions of the rat forebrain.<br />

C. Generation of the ER null mice<br />

The field of estrogen action has been and continues to be<br />

a dynamic one with broadening scope. Although estradiol is<br />

often called the female sex steroid, the list of tissues and<br />

organs in which estrogen actions appear to be critical to<br />

physiological function continues to expand, in both sexes.<br />

The crucial actions of estradiol in the female reproductive<br />

tract, breast, and hypothalamic-pituitary axis have been well<br />

described. However, we now know of significant functions<br />

fulfilled by estrogens in male reproduction. Furthermore, the<br />

effects of estrogens in the physiology, maintenance, and<br />

overall health of the cardiovascular, central nervous, bone,<br />

immune, skin, and adipogenesis systems have been an area<br />

of continued research. And finally, the role that estrogen and<br />

normal and variant ERs may play in carcinogenesis, especially<br />

in the breast and female reproductive tract, continues<br />

to receive great attention.<br />

The rationale for generating mice that possess no functional<br />

ER is multifaceted, but in its most simple terms, was<br />

founded on the classical ablation experiments of the early<br />

part of this century. In 1900, Knauer (105) described the<br />

ability of ovarian grafts to prevent uterine atrophy in the<br />

castrated guinea pig. Five years later, Marshall and Jolly (106)<br />

described the capacity of ovarian extracts to induce estrus<br />

when administered to ovariectomized dogs. Similar protocols<br />

were later elegantly employed by Jost (107) to substantiate<br />

the endocrine function of the testis and the importance<br />

of testosterone in sex determination. These studies were critical<br />

to establishing the following basic criteria required to<br />

verify an endocrine role for a particular organ or tissue: 1)<br />

removal or destruction of the synthesizing organ should<br />

result in predictable symptoms presumed to be related to the<br />

absence of the hormone; 2) administration of material prepared<br />

from the removed organ should relieve these symptoms;<br />

and 3) the hormone should be present in and extractable<br />

from both the organ and blood (108). In the spirit of these<br />

earlier studies, the later part of this century has witnessed the<br />

marriage of two relatively new methodologies to introduce<br />

a modern version of the ablation experiment. The combination<br />

of in vitro culture of mouse embryonic stem cells and<br />

targeted homologous recombination has generated a tool<br />

that allows for the precise disruption or knockout of a particular<br />

gene of study and the passage of this mutation to<br />

offspring. Although one can debate whether this new technique<br />

is more or less invasive than the previous surgical<br />

methods, it is obvious that the classic “ablation” experiment<br />

has been elevated to a molecular level. The function of a<br />

specific hormone can now be studied rather than the function<br />

of a whole endocrine organ that may produce multiple secretions.<br />

Furthermore, this new technology allows for the<br />

study of a particular cellular component, such as a receptor,<br />

that is intrinsic to one or more tissues. Such studies were<br />

previously impossible or relied on the use of chemical antagonists,<br />

which introduced their own inherent limitations.<br />

Additionally, gene targeting provides for in vivo methods to<br />

study the roles of a particular receptor throughout the life of<br />

the animal, including the early development stages. The current<br />

tools that allow for the generation of transgenic mice<br />

have already made significant contributions to our knowledge<br />

of particular genes, especially those involved in development<br />

and reproduction (reviewed in Refs. 109 and 110).<br />

At the time the ERKO mouse was envisioned, the ER� was<br />

the only form of ER known to exist. Furthermore, there were<br />

no reports of ER mutations in normal tissue that resulted in<br />

estrogen insensitivity in humans or laboratory animals. This<br />

was in contrast to the descriptions of syndromes of receptorbased<br />

insensitivity to androgens (37, 38) and thyroid hormones<br />

(40). Therefore, investigators were inclined to conclude<br />

that mutations that resulted in insufficient estrogen<br />

synthesis or resistance to the hormone at the level of the<br />

target organ were lethal at the earliest developmental stages<br />

(111, 112). This view was strengthened by reports of the<br />

detection of ER� mRNA in both human (113) and mouse<br />

(114) oocytes as well as in mouse blastocysts (115). Accordingly,<br />

the concept of generating a mouse devoid of ER� was<br />

initially met with skepticism but was pursued as a collaborative<br />

effort between our laboratory and that of Dr. Oliver<br />

Smithies of the University of North Carolina at Chapel Hill.<br />

If disruption of the mouse ER� gene did prove lethal, a model<br />

to study the precise time and locations of critical ER-mediated<br />

actions during early development would be available.<br />

However, providing the animal was viable, an in vivo model<br />

of estrogen insensitivity would now be accessible for continued<br />

study. Six years after the initial description of the<br />

�ERKO mouse, we now take for granted that disruption of<br />

the ER� gene proved not to be lethal, but rather the animal<br />

develops normally and exhibits a life span comparable to its<br />

wild-type litter mates (46). However, as will be elaborated in<br />

detail in this review, the adult �ERKO mice exhibit several<br />

abnormalities and deficiencies, most notable of which are the<br />

phenotypic syndromes that result in infertility in both sexes.<br />

Since this time, additional discoveries have been made to<br />

enhance the utility of the �ERKO model. Soon after the<br />

generation of the �ERKO mouse, Smith et al. (116) reported<br />

the first and only known case of clinical estrogen insensitivity<br />

due to an inactivating mutation of the ER� gene in a human<br />

patient. Later came multiple descriptions of aromatase and<br />

subsequent estradiol deficiency in humans (117–120). And<br />

finally, a second collaborative effort resulted in the successful<br />

generation of the �ERKO mice, which also survive to adulthood<br />

and exhibit phenotypes unique from those of the<br />

�ERKO (47). However, splicing variants of the disrupted<br />

genes that may encode for receptors with decreased functional<br />

activity have been detected in small amounts in each<br />

of the respective ERKO models (as discussed below) and<br />

therefore complicate the interpretation of the nonlethality of<br />

the gene targeting. Nonetheless, we now know that a loss of<br />

full function of any one of the two ERs is neither lethal nor<br />

detrimental to embryonic and fetal development in both<br />

mice and humans. Perhaps the survival of the ERKO mice,<br />

the aromatase-deficient humans, and the ER�-deficient male<br />

will prompt a renewed effort among clinicians to suspect and<br />

investigate the possibility of estrogen insufficiency or resistance<br />

in patients not responding to conventional therapies.<br />

The ERKO mice provide broad and multiple advantages<br />

to the research efforts toward understanding the function


366 COUSE AND KORACH Vol. 20, No. 3<br />

and mechanism of estogen action. Much of what is known<br />

about estrogen action was inferred from in vivo studies involving<br />

castration or the administration of ER antagonists or<br />

inhibitors of estradiol synthesis. These findings have been<br />

complemented by the vast knowledge gained from in vitro<br />

cell culture studies, employing chimeric and mutant versions<br />

of the ER, varied cell types, multiple combinations of promoter-reporter<br />

gene constructs, as well as synthetic agonists<br />

and antagonists. However, there are distinct disadvantages<br />

to these experimental schemes. Studies using aromatase inhibitors<br />

and/or estrogen antagonists are complicated by several<br />

factors, including variability of the compound to block<br />

the action of the natural hormone or enzyme. The effectiveness<br />

of various antagonists is highly dependent upon the<br />

animal model, the tissue or cell of study, the bioavailability<br />

of the compound at different target tissues, and the class of<br />

antiestrogen used (8). This dilemma is further complicated<br />

by the discovery of the ER�, since no known ER-selective<br />

agonists or antagonists have been characterized at an in vivo<br />

level. The limitations of in vitro cell culture experimental<br />

approaches are obvious and mostly based on their finite<br />

application to the whole animal. Therefore, the ERKO models<br />

provide a unique tool to investigate the role of the ER in<br />

the context of the whole animal, and equally important,<br />

during the complete life span of the animal. At their most<br />

fundamental level, the ERKO mice address the role of the ER<br />

in the development and normal physiology of all organ systems,<br />

as well as in carcinogenesis, toxicity, and aging. Furthermore,<br />

unlike the “castrate” model in which several hormones<br />

are removed from the system, the ERKO mice retain<br />

the capability to synthesize the gonadal steroids, including<br />

the natural ER ligand, estradiol. Therefore, the biochemical<br />

functions of estradiol and the ER can be investigated in the<br />

presence of presumably intact pathways for the other gonadal<br />

hormones. The presence of estrogens in the absence of<br />

ER also provides for the possibility of discovering pathways<br />

of estrogen action that are independent of nuclear ER, or<br />

mediated via previously unknown forms of the receptor.<br />

Additionally, although the majority of in vitro studies indicate<br />

that ER� and ER� may have redundant functions, their<br />

differences in tissue distribution and response to certain ligands<br />

indicate the presence of distinct roles fulfilled by each.<br />

The fact that the �ERKO mice exhibit an unaltered pattern of<br />

ER� mRNA expression strengthens the usefulness of this<br />

model to dissect these potential ER�-mediated actions (93,<br />

121). Finally, consistent with those criteria discussed earlier<br />

for establishing an endocrine function to an organ, the ERKO<br />

animals now provide a null background available for transgenic<br />

reintroduction of the ER of other species, mutated ERs,<br />

or targeted ER expression to a specific tissue or cell type.<br />

A detailed description of the targeting scheme employed<br />

to disrupt the mouse ER� gene can be found in the initial<br />

description of the �ERKO mice (46). As shown in Fig. 1, a<br />

1.8-kb insert possessing the gene for neomycin (neo) resistance<br />

under the control of the phosphoglycerate kinase<br />

(PGK) promoter and including a PGK-polyadenylation signal<br />

was inserted into a NotI site in exon 2 of an ER� gene<br />

fragment subcloned from a genomic library of 129/J mouse<br />

DNA. The targeting insert was placed in a replacement �<br />

type targeting vector (122) with the appropriate ER� gene-<br />

flanking sequences. Upon successful targeting in mouse embryonic<br />

stem cells (129/J), the neo insert is placed approximately<br />

270 bp downstream of the ER� translation start site<br />

and thereby inhibits proper expression of the ER� gene. Since<br />

this was the current state of the technology, no portion of the<br />

ER� gene was removed during the targeting event. Standard<br />

protocols of clone selection and blastocyst (C57BL/6J) injection<br />

were used to generate chimeric mice possessing the<br />

disrupted gene, some of which demonstrated germ-line<br />

transmission of the mutation when bred with wild-type<br />

mates (122). Southern blot and PCR analysis of genomic<br />

DNA from mice of all three genotypes indicated the correct<br />

targeting of the ER� gene and the absence of any heterologous<br />

recombination in other regions of the genome. Inbreeding<br />

of mice heterozygous for the ER� disruption resulted in<br />

a Mendelian distribution of all three genotypes as well as a<br />

balanced sex ratio, indicating that the ER� is not critical to sex<br />

determination at the level of the external genitalia (46).<br />

The generation of mice homozygous for a disruption of the<br />

ER� gene was similar to that described above for the �ERKO<br />

and can be found in detail in the initial description (47). A<br />

genomic clone that spanned a 15-kb region possessing the<br />

first three exons of the mouse ER� gene was selected from<br />

a 129/SvJ mouse library. A replacement � type targeting<br />

construct was generated to include 5� and 3� homologous<br />

sequences of 1.3 and 7.4 kb, respectively (Fig. 1). A PGK<br />

promoter-regulated neo gene was inserted in the reverse<br />

orientation into a PstI site in exon 3 of the ER� clone. Therefore,<br />

correct targeting resulted in disruption of the sequences<br />

coding for the first zinc finger of the ER� protein, a domain<br />

critical to normal function of the receptor. Chimeric and<br />

heterozygous offspring were generated as described above<br />

for the �ERKO mice. Once again, mice possessing the targeted<br />

disruption of the ER� gene were identified by diagnostic<br />

Southern blotting and PCR of genomic DNA. As with<br />

the �ERKO, inbreeding of mice heterozygous for the disruption<br />

yielded a Mendelian distribution of all three genotypes<br />

as well as a balanced sex ratio (47).<br />

In both knockout models, RT-PCR on RNA from target<br />

tissues indicated the presence of multiple splicing variants of<br />

the respective ER transcripts (47, 123). In neither case has<br />

wild-type-like mRNA transcribed from the disrupted receptor<br />

gene been detected. The greater proportion of the variants<br />

detected in each ERKO model possessed frame shifts that<br />

would result in a severely truncated or mutated ER if translated.<br />

However, in the �ERKO, a single-splice variant capable<br />

of encoding a mutant ER� protein with significantly<br />

decreased transactivational capacity in vitro was detected at<br />

very low levels (123). A similar ER� splice variant, in which<br />

the reading frame was preserved although coding sequences<br />

were removed, was detected in ovaries of the �ERKO mice.<br />

This single variant, if translated, would encode a mutant ER�<br />

lacking the first zinc finger and therefore would be unlikely<br />

to transactivate due to an inability to tightly associate with<br />

the chromatin structure within the regulatory regions of target<br />

genes.<br />

This is not the first report of targeted insertions resulting<br />

in aberrant splicing of a disrupted gene. <strong>Mice</strong> possessing a<br />

targeted disruption of the transforming growth factor-� gene<br />

produce a transcript in which the entire exon possessing the


June, 1999 ESTROGEN RECEPTOR NULL MICE 367<br />

disrupting insert is accurately removed via the conventional<br />

donor and acceptor splice sites, preserving the normal reading<br />

frame of the gene (124). Studies in three different human<br />

genes have demonstrated that point mutations resulting in a<br />

premature stop codon can lead to complete excision of the<br />

exon possessing the mutation (125, 126). Furthermore, Reed<br />

and Maniatis (127) used artificial deletions and insertions<br />

within exons of genes that normally display alternative splicing<br />

to demonstrate that the proximity of the acceptor and<br />

donor splicing sequences to one another plays a role in splicing<br />

mechanisms. It is possible that insertion of sequences as<br />

large as those used in the ERKO mice may disrupt the spatial<br />

requirements necessary for proper mRNA splicing. Therefore,<br />

the above studies, along with our experiences, are relevant<br />

to the practice of targeting genes by insertion of a large<br />

disrupting sequence possessing internal stop codons.<br />

1. Interpretation of phenotypes in receptor null mice. The use of<br />

methodologies to target and disrupt individual genes has<br />

created numerous models available for study (reviewed in<br />

Refs. 109, 110, and 128). Furthermore, this impact has been<br />

felt in several facets of the biological sciences. Although it<br />

may be initially thought that a particular gene plays no role<br />

in the physiology of certain animal systems, such as reproduction<br />

or behavior, disruption of the gene and the subsequent<br />

phenotypes often prove otherwise. Therefore, transgenic<br />

and knockout technologies have spawned a number of<br />

collaborative efforts between investigators of varied disciplines<br />

that may have never occurred.<br />

An issue that has become apparent from the numerous<br />

gene disruption studies and interdisciplinary collaborative<br />

efforts is a collection of caveats to be considered when evaluating<br />

phenotypical data from a knockout model. These have<br />

arisen mostly from the behavioral sciences (reviewed in Refs.<br />

129 and 130), but have expanding application to all areas of<br />

study provided by transgenic animals. The first of these<br />

caveats is one that may be most relevant to the steroid receptor<br />

mutant models, i.e., when studying the target tissue of<br />

an adult receptor-knockout mouse, one must realize the tissue<br />

passed through all the stages of development and “organization”<br />

in the absence of the respective receptor. Therefore,<br />

this tissue, and in essence the whole animal, cannot be<br />

assumed to be truly identical to the wild type in all aspects<br />

except for the absence of the targeted gene product. Any<br />

genetic redundancies or compensatory mechanisms that<br />

took place during development cannot be readily detected or<br />

accounted for during most experiments. Therefore, the lack<br />

of a phenotype does not necessarily discount the function of<br />

the disrupted gene in the physiology being studied. Additionally,<br />

it is difficult to distinguish between an organizational<br />

vs. activational basis for an observed deficit in a particular<br />

physiology when studying the adult mutant. For<br />

example, observed resistance to a hormone due to alterations<br />

downstream of the function of the disrupted gene may be<br />

apparent during adulthood, but may have been imprinted<br />

during development.<br />

Other caveats of interpreting data from receptor null mice<br />

are founded in the methods used to generate and maintain<br />

a line of knockout mice. The standard protocol for generating<br />

a knockout mouse involves the incorporation of embryonic<br />

stem cells of a 129 strain of mice that carry the desired<br />

mutation into the blastocyst of a C57BL/6 strain. The resulting<br />

chimeric animals are then often back-crossed to the<br />

C57BL/6 strain once again until mice homozygous for the<br />

disruption are acquired. Therefore, early generations of<br />

knockout mice are composed of a somewhat chimeric genome,<br />

especially in the chromosomal regions closest to the<br />

targeted locus. This is of special importance in behavioral<br />

studies in light of the known variations in the sexual behavior<br />

of different strains of mice (131). However, most relevant to<br />

the ERKO, significant variations in estrogen responsiveness<br />

of the female reproductive tract among the different strains<br />

of mice are also known to exist (132, 133). A recent report by<br />

Roper et al. (134) has further defined the genetic basis for the<br />

variations that exist in the effects of estradiol on classical<br />

uterine parameters in mice. These limitations can be overcome<br />

to some degree by increasing experimental sample<br />

sizes and including parental strains as a control group in all<br />

experiments (135). In addition, the various models of hormone<br />

and steroid receptor deficiency that are now available<br />

no longer make necessary the complete interpretation of data<br />

from any one model. Therefore the data from these models,<br />

when interpreted as a whole, should prove invaluable to<br />

elucidating the roles of the different sex steroid receptors in<br />

both development and adult physiology.<br />

III. Reproductive Tract Phenotypes of the Female<br />

The most well characterized estrogen target tissues are<br />

those of the mammalian female reproductive tract, comprised<br />

of the ovaries, oviducts, uterus, cervix, and vagina.<br />

Reproductive capabilities in the female are dependent on the<br />

sequential processes of differentiation during the embryonic<br />

and prenatal periods and the maturation of these tissues<br />

during puberty. Differentiation of the fetal gonads to ovaries<br />

results from a lack of the Y chromosome, and therefore the<br />

testis-determining genes, and the presence of putative, yet<br />

unidentified, autosomal ovary-determining genes (111). The<br />

ductal organs of the tract subsequently result from differentiation<br />

of the fetal ambisexual precursor, the Müllerian<br />

ducts, due to the lack of testicular hormones. Early studies<br />

indicated that the female reproductive tract is the default<br />

phenotypical sex and will differentiate and develop normally<br />

in the absence of ovaries and adrenal glands (107, 136). Therefore,<br />

it appears that estrogens are not required for differentiation<br />

and initial development of the female reproductive<br />

tract, whereas testosterone is critical to differentiation of the<br />

male genitalia. This conclusion has been reconfirmed by the<br />

male-to-female sex reversal observed in male mice homozygous<br />

for a targeted disruption of the gene encoding SF-1, a<br />

transcription factor that regulates the steroid hydroxylases in<br />

steroidogenic cells (137, 138). Despite an inability to synthesize<br />

steroids and a complete lack of gonads, all SF-1 knockout<br />

mice develop female internal genitalia, regardless of genetic<br />

sex (137). It is therefore not surprising that the �ERKO and<br />

�ERKO female mice exhibit a properly differentiated female<br />

reproductive tract possessing the constituent structures (46,<br />

47). However, estrogen insensitivity has severely disrupted<br />

sexual maturation of the whole reproductive tract in the


368 COUSE AND KORACH Vol. 20, No. 3<br />

�ERKO female and ovarian function in the �ERKO female.<br />

The consequences of ER gene disruption on the individual<br />

components of the female reproductive tract is the topic of<br />

this portion of the review.<br />

Before we continue, we believe it is necessary to briefly<br />

reiterate those studies carried out to verify successful targeting<br />

of the ER� gene in the �ERKO. This discussion is<br />

appropriate for this portion of the review because the majority<br />

of these experiments were performed on uterine tissue.<br />

To determine the effectiveness of the gene targeting, Western<br />

blots of adult �ERKO uterine nuclear and cytosolic extracts<br />

were probed with the H222 antibodies, a rat monoclonal<br />

antibody specific to the ligand-binding domain of the human<br />

ER� (10). Our studies, as well as those of others, have demonstrated<br />

that this antibody possesses high cross-reactivity to<br />

the mouse ER� (139–141). These assays detected no wildtype<br />

ER� or any other immunoreactive fragments unique to<br />

the �ERKO uterus. Similar results were obtained when blots<br />

were probed with the rabbit antiserum ER-21, directed toward<br />

the 21 amino-terminal residues of the rat ER� (141).<br />

However, binding assays using 3 H-E 2 on �ERKO uterine<br />

extracts indicated the presence of high-affinity binding of the<br />

hormone at levels approximately 3–9% of the wild type (123).<br />

In agreement with these data, sucrose gradient analysis with<br />

3 H-E2 on low-salt cytosol extracts from �ERKO uteri indicated<br />

a binding factor with an 8S sedimentation value, similar<br />

to that of the wild-type ER� (123). The discovery of the<br />

ER�, reported approximately 3 yr after the generation of the<br />

�ERKO, prompted a renewed assessment of this �ERKO<br />

estrogen-binding data in several publications. Unfortunately,<br />

in a number of these reports, the original datum<br />

discussed above is not evaluated in full, and the authors<br />

elude to ER� as the likely binding source in the �ERKO uteri.<br />

Certainly at the time of the initial characterization, concern<br />

over the residual level of binding in the �ERKO uteri was<br />

often mixed with the wonder of possibly discovering an<br />

unknown ER. However, during these studies we also demonstrated<br />

that when the H222 antibodies were included in<br />

the sucrose gradient assays, the estradiol binding peak in the<br />

�ERKO uterine extract was shifted accordingly (123). The<br />

H222 antibodies have been shown by us, as well as by other<br />

laboratories, to be ER� specific and unable to recognize ER�<br />

by Western blot analysis or immunohistochemistry (142). As<br />

described earlier in this review, our RT-PCR analysis on<br />

mRNA from �ERKO uteri demonstrated the presence of a<br />

splicing variant of the disrupted ER� gene that could encode<br />

a mutant ER� possessing both the ability to bind estradiol as<br />

well the H222 epitope (123). Furthermore, we have recently<br />

shown that ER� mRNA is undetectable in the uteri of adult<br />

wild-type as well as �ERKO mice when assayed by ribonuclease<br />

protection assay (93). Therefore, we believe that relatively<br />

conclusive data have been generated to indicate that<br />

the estradiol-binding factor present in �ERKO uteri is most<br />

likely not ER�.<br />

A. Uterus<br />

1. Uterine phenotype and estrogen insensitivity. The ER has been<br />

detected by steroid autoradiography and immunohistochemical<br />

methods in the ductal structures of the rodent fe-<br />

male reproductive tract during several stages of development,<br />

including the late fetal and neonatal stages through<br />

puberty and adulthood (reviewed in Refs. 112 and 143).<br />

Several reports describe the initial appearance of ER immunoreactivity<br />

in the developing uterus as early as fetal day 15<br />

(112, 143). ER immunoreactivity was first detectable in mesenchymal<br />

cells, whereas induction in the epithelial cells occurs<br />

during the late fetal stages and increases significantly<br />

during the neonatal period (112, 143). The fully developed<br />

uterus is composed of many heterogeneous cell types comprising<br />

three major anatomical compartments, the outer<br />

myometrium, endometrial stroma, and luminal/glandular<br />

epithelium. In the immature CD-1 mouse, ER� immunoreactivity<br />

is easily detectable in the stroma on day 1 and continues<br />

to rise to a maximal level on day 10, whereas the<br />

appearance of epithelial ER� is delayed and reaches a peak<br />

around day 16 (144). Other reports indicate variations in the<br />

exact timing of the appearance of ER� among different<br />

strains and species, most likely reflecting temporal deviations<br />

in development (112, 143).<br />

The presence of an intact estrogen-signaling system appears<br />

to coincide with the appearance of ER�. In several<br />

species, estrogen treatment of fetal and neonatal females<br />

results in the stimulation of increased uterine levels of nucleic<br />

acid (136, 145), protein synthesis (146), ornithine decarboxylase<br />

(147), progesterone receptor (148), and cellular<br />

proliferation (145, 149, 150). However, a full biological response<br />

to estradiol in terms of maximum increases in uterine<br />

weight is not possible in the neonatal uterus, and can be<br />

observed only after the animal approaches weaning age<br />

(146). Furthermore, significant differences in the uterine response<br />

to estradiol between the neonate and sexually mature<br />

rodent are known (151). For example, estrogen stimulates<br />

cellular proliferation in all tissues of the immature uterus,<br />

whereas this response becomes limited to the epithelial compartment<br />

during adulthood (151, 152). Therefore, sexual maturation<br />

of the uterus is not simply marked by the presence<br />

of ER�, but rather the acquisition of the capacity to undergo<br />

the correct synchronized phases of proliferation and differentiation<br />

elicited by the ovary-derived sex steroids.<br />

As shown in Fig. 3, the uteri of both adult �ERKO and<br />

�ERKO females possess all three definitive uterine compartments,<br />

the myometrium, endometrial stroma, and epithelium.<br />

However, in the �ERKO, each is hypoplastic and results<br />

in whole uterine weights that are approximately half<br />

that recorded for wild-type littermates (46). In contrast, the<br />

uteri of adult �ERKO females appear normal and able to<br />

undergo the cyclic changes associated with the ovarian steroid<br />

hormones (47). Therefore, perinatal development of the<br />

female reproductive tract in the mouse appears to be independent<br />

of ER� and ER� actions. However, estrogen responsiveness<br />

and subsequent sexual maturity in the uterus has<br />

been ablated by disruption of the ER� gene. The �ERKO<br />

endometrial stoma is characterized by a less organized structure<br />

and hypotrophy, with a sparse distribution of uterine<br />

glands compared with that of the wild type (153). Luminal<br />

and glandular epithelial cells in the �ERKO uterus most often<br />

appear healthy, but are consistently cuboidal and lack the<br />

normal “estrogenized” morphology of a tall columnar shape<br />

and basally located nucleus (Fig. 3). This phenotype is in-


June, 1999 ESTROGEN RECEPTOR NULL MICE 369<br />

FIG. 3. Histology of uterine and vaginal tissue of wild-type, �ERKO, and �ERKO females. Cross-sections of uterine tissue from (a) wild-type,<br />

(b) �ERKO, and (c) �ERKO adult females illustrating the presence of all three anatomical tissue compartments in the uteri of the wild-type<br />

and ERKO mice (33�; inset, 132�). The representative wild-type uterine section illustrates a normal myometrium (My), endometrial stroma<br />

(St), and epithelium (Ep) (inset). The representative �ERKO uterine section illustrates the characteristic hypoplasia of each compartment, a<br />

slightly disorganized endometrial stroma, and a lack of estrogenization of the luminal and glandular epithelium (inset). Note the dramatically<br />

smaller diameter of the �ERKO uterus, as indicated by the ability to fit the whole transverse section of the tissue in the field of view. The<br />

representative �ERKO uterine section is indistinguishable from that of the wild-type, including the presence of estrogen-stimulated luminal<br />

epithelium (inset). Below are cross-sections of vaginal tissue from representative (d) wild-type, (e) �ERKO, and (f) �ERKO female mice (66�).<br />

The representative wild-type vaginal section illustrates a normal stroma (St) and hypertrophied epithelium (Ep) showing estrogen-induced<br />

stratification and cornification. In contrast, these estrogen actions are lacking in the �ERKO vagina. Once again, the tissue of the �ERKO is<br />

indistinguishable from the wild-type control. Scale bar � 1 �m.<br />

teresting in light of the increased levels of estradiol found in<br />

the serum of adult �ERKO females (Table 2). However,<br />

Lindzey et al. (154) demonstrated that the concurrently elevated<br />

ovary-derived androgens in the �ERKO female (Table<br />

2) do provide for some maintenance of uterine weight, which<br />

can be further reduced upon ovariectomy. We recently reported<br />

that ER� mRNA is barely detectable in the adult<br />

mouse uterus, including those from �ERKO mice (93), making<br />

it unlikely that ER� could provide a compensatory role<br />

in the �ERKO uterus. Numerous immunohistochemical<br />

studies for ER� and the apparent loss of estrogen sensitivity<br />

in the �ERKO uterus indicate that the classical ER� is the<br />

predominant form responsible for mediating estrogen actions<br />

in the mouse uterus.<br />

The response of the ovariectomized rodent uterus to estradiol<br />

has been well documented and is often described as<br />

biphasic, with the initial phase consisting of effects that become<br />

apparent within the first 6hofasingle estrogen treatment<br />

(reviewed in Ref. 155). These include metabolic re-<br />

sponses in the form of increased water imbibition, vascular<br />

permeability and hyperemia, prostaglandin release, glucose<br />

metabolism, and eosinophil infiltration (155). A series of<br />

biosynthetic responses are also characteristic of the first<br />

phase and include increased RNA polymerase and chromatin<br />

activity, lipid and protein synthesis, and increased glucose-6-phosphate<br />

dehydrogenase (155). As several of the<br />

above processes continue, they are accompanied by dramatic<br />

increases in RNA and DNA synthesis, mitosis, and cellular<br />

hyperplasia and hypertrophy that peak 24–72 h after exposure<br />

(155). Initial studies suggested that this biphasic pattern<br />

may be the result of at least two types of acceptor sites for<br />

estradiol-receptor complexes in the uterus (156). However,<br />

the uteri of �ERKO females fail to exhibit components of both<br />

phases after estrogen treatment, providing strong evidence<br />

for the requirement of ER� in the full response (46, 123). In<br />

brief, when treated with 40 �g E 2 or diethylstilbestrol (DES)<br />

per kg body weight for three consecutive days, wild-type<br />

mice exhibited the expected 3- to 4-fold increase in uterine


370 COUSE AND KORACH Vol. 20, No. 3<br />

TABLE 2. Serum hormone levels in adult wild-type and �ERKO mice<br />

Hormone<br />

Gonadal steroids<br />

Wild-type (SEM)<br />

Female<br />

�ERKO (SEM) Wild-type (SEM)<br />

Male<br />

�ERKO (SEM)<br />

Estradiol (pg/ml) b<br />

29.5 � 2.5 84.3 � 12.5 a<br />

11.8 � 3.4 12.9 � 3.4<br />

Progesterone (ng/ml) b<br />

2.3 � 0.6 4.0 � 1.1 0.5 � 0.3 0.3 � 0.1<br />

Testosterone (ng/ml)<br />

Anterior pituitary<br />

0.4 � 0.4 3.2 � 0.6 9.3 � 4.0 16.0 � 2.3<br />

LH (ng/ml) 0.3 � 0.04 1.7 � 0.3 a<br />

2.4 � 1.2 3.7 � 0.7<br />

FSH (ng/ml) 4.9 � 0.6 5.4 � 0.7 26.0 � 1.4 30.0 � 1.1<br />

PRL (ng/ml)<br />

nd, Not determined.<br />

a<br />

t test, wild-type vs. ERKO, P � 0.001.<br />

18.8 � 10.7 3.5 � 1.3 nd nd<br />

b<br />

These values in the female are different than those reported in Ref. 123, which were carried out on pooled sera. The values above are the<br />

means from assays on individual samples and therefore are more likely to reflect the true levels in the two genotypes.<br />

wet weight, whereas no such response was observed in the<br />

uteri of �ERKO mice (46, 157). It should be noted that this<br />

pharmacological dose of estrogen is well beyond that required<br />

to achieve a maximum response in the wild-type<br />

rodent. Nonetheless, estrogen-treated �ERKO uteri exhibited<br />

no apparent components of the initial phase of estrogen<br />

effects, including water imbibition and hyperemia. Histological<br />

analysis and [ 3 H]thymidine incorporation assays indicated<br />

a lack of significant cellular proliferation and DNA<br />

synthesis in uteri from the estrogen-treated �ERKO mice<br />

(123, 153). Interestingly, although the heterozygous females<br />

possess approximately one-half the normal complement of<br />

ER�, their uterine response to estrogens is equal to that of the<br />

wild-type females. In a similar study, wild-type and �ERKO<br />

mice treated with hydroxy-tamoxifen (1 mg/kg) produced<br />

comparable results (157), eliciting the expected estrogenic<br />

response in the wild-type and having no effect on the �ERKO<br />

uterus. These studies thereby confirm that the estrogen agonist<br />

activity of hydroxy-tamoxifen, which is somewhat<br />

unique to the mouse uterus (8), is mediated via the ER�<br />

pathway.<br />

The mitogenic and stimulatory action of estradiol in the<br />

uterus is a complex process involving increased RNA polymerase<br />

and ribosomal activity (158), resulting in the regulation<br />

of a plethora of genes. It is well accepted that the<br />

ligand-bound ER complex is not directly involved in the<br />

mediation of all responses elicited by estrogens in the uterus,<br />

but rather serves as a stimulus for a cascade of signaling<br />

pathways that act to amplify the estrogen action. However,<br />

certain genes appear to be directly regulated by the ER�estradiol<br />

complex and possess functional estrogen-responsive<br />

elements within their regulatory regions. Two such examples<br />

are the genes encoding the progesterone receptor<br />

(PR) (159, 160) and the secretory protein, lactoferrin (161). In<br />

fact, the regulation of the uterine PR and lactoferrin genes<br />

have often been used as assays for the estrogenic activity of<br />

experimental compounds. Therefore, with a similar intent,<br />

we used these estrogen markers to attest for estrogen insensitivity<br />

in the uteri of the �ERKO mouse. A single dose of<br />

estradiol, known to be effective in inducing the PR and lactoferrin<br />

genes within 24 h in uteri of wild-type mice, produced<br />

no such up-regulation in the uteri of the �ERKO mice,<br />

confirming the need for a direct action of the ER� in this<br />

mechanism (123). Interestingly, a recent report by Tibbetts et<br />

al. (162) demonstrated that the estrogen-stimulated increases<br />

in PR are localized to the stromal and myometrial compartments,<br />

whereas the increases in lactoferrin are isolated to the<br />

luminal and glandular epithelium in the mouse uterus.<br />

Therefore, disruption of the ER� gene has resulted in estrogen<br />

insensitivity in all three anatomical compartments of the<br />

uterus. However, it must be noted that constitutive levels of<br />

PR and lactoferrin mRNA are present in the �ERKO uteri,<br />

suggesting that these genes are also under the influence of<br />

pathways independent of ER�. A testimony to the complexity<br />

of estrogen action in the uterus is the finding that while<br />

estradiol up-regulates PR expression in the myometrium and<br />

stroma, it simultaneously abolishes PR levels in the luminal<br />

epithelium (162). This would indicate an inhibitory role of the<br />

estradiol-ER� complex on PR expression in this portion of the<br />

uterus. Speculating that this pathway may therefore be lacking<br />

in the �ERKO uterus, an investigation as to the source of<br />

the PR mRNA in the �ERKO uteri is warranted.<br />

2. Changes in growth factor functions. A component of the<br />

cascade of events that lead to the obvious changes in the<br />

physiology of the adult uterus after estrogen exposure are the<br />

auto- and paracrine actions of polypeptide growth factors.<br />

Several members of the epidermal growth factor family have<br />

been suggested as possible mediators of estrogen-induced<br />

mitogenesis in the uterus. This hypothesis is based on experiments<br />

demonstrating that estradiol up-regulates the<br />

uterine levels of epidermal growth factor and its receptor<br />

(EGF, EGF-R) (163, 164), transforming growth factor-� (165),<br />

and insulin-like growth factor-I (IGF-1) (166). Furthermore,<br />

mice homozygous for a targeted disruption of the EGF-R<br />

gene exhibit a hypoplastic uterus that is significantly reduced<br />

in size (167), similar to that of the �ERKO. Experimental data<br />

indicate that treatment of ovariectomized wild-type mice<br />

with EGF mimics the early effects of estradiol and DES in<br />

terms of inducing modified cell morphology and increases in<br />

the levels of ER, DNA synthesis, phosphatidylinositol turnover,<br />

PR, and lactoferrin in the uterus (168–170). Further<br />

studies have illustrated that cotreatment with anti-EGF antibodies<br />

was able to attenuate the uterine response to estradiol,<br />

presumably due to inactivation of the EGF-signaling<br />

pathway (168). In turn, cotreatment with the estrogen antagonist<br />

ICI-164,384 was able to reduce the uterine response<br />

to EGF (169). These in vivo studies suggest a cross-talk mechanism<br />

between the EGF and ligand-independent ER- signaling<br />

pathways. The results of the animal studies have been


June, 1999 ESTROGEN RECEPTOR NULL MICE 371<br />

supported by numerous in vitro experiments demonstrating<br />

ligand-independent activation of the nuclear signaling pathway<br />

of the ER�, possibly via altering the phosphorylation<br />

pattern of the ER� (reviewed in Ref. 55). The culmination of<br />

these and several other studies has led to the proposed model<br />

in which the mitogenic actions of estradiol in the rodent<br />

uterus appear to be at least partially mediated by EGF; however,<br />

in turn the mitogenic effects of EGF require the presence<br />

of ER�.<br />

Therefore, the �ERKO female provides an excellent in vivo<br />

model to study this cross-talk between the ER�- and EGFsignaling<br />

systems in the uterus. The uteri of �ERKO females<br />

possess wild-type levels of functional EGF and EGF-R (170).<br />

Nonetheless, the mitogenic actions and induction of estrogen-responsive<br />

genes elicited by EGF in the wild-type uterus<br />

have been ablated in the �ERKO, confirming the interaction<br />

of these two signaling systems (170). However, not all EGF<br />

responses are lacking in the uteri of �ERKO females, as this<br />

same study demonstrated that the mechanisms for EGFmediated<br />

up-regulation of the c-fos gene remained intact<br />

(170). These studies have thereby confirmed the need for<br />

functional ER� for the mitogenic actions of EGF in the uterus.<br />

Cunha et al. has extended the use of the �ERKO mouse to<br />

investigate the intersecting roles of ER�-mediated estrogen<br />

stimulation and growth factors in the uterus through a series<br />

of tissue recombination experiments. The observation of estrogenic<br />

effects in wild-type uterine epithelial cells that are<br />

apparently lacking ER� has prompted numerous investigations<br />

to illustrate a role for paracrine factors secreted by the<br />

underlying ER�-positive stromal compartment, and thereby<br />

mediating the epithelial response (143). These studies have<br />

been advanced by methods that provide for the delicate<br />

construction of tissue recombinants, in which uterine stoma<br />

and epithelium are enzymatically disassociated and recombined<br />

with similar tissue from uteri from animals of different<br />

treatments or models to ultimately regenerate a chimeric<br />

stromal-epithelial unit (reviewed in Ref. 143). These tissue<br />

recombinants are implanted under the kidney capsule of<br />

ovariectomized nude mice, which are then acutely treated<br />

with estrogen agonists or antagonists. Later removal of the<br />

recombinant grafts allows for the evaluation of certain end<br />

points of estrogen action in each portion of the recombinant.<br />

Cooke et al. (171) described experiments in which wild-type<br />

(ER��) uterine stroma were recombined with �ERKO<br />

(ER��) uterine epithelium and vice versa. The results of these<br />

studies illustrate that proliferation of the epithelial portion of<br />

the recombinant was possible only when ER�� stroma were<br />

present and did not require ER� in the epithelium (171).<br />

Interestingly, similar recombinant experiments using tissue<br />

from the EGF-R knockout mice illustrated that the estrogensignaling<br />

pathways required for stimulation of the stroma<br />

and subsequent induction of epithelial growth are intact in<br />

the absence of EGF-R (167). Aside from the proliferative<br />

effects of estradiol, previous studies suggested that estrogen<br />

stimulation of secretory products from uterine epithelium,<br />

e.g., lactoferrin, is directly mediated by the epithelial ER�<br />

(140, 162). However, Buchanan et al. (172) recently employed<br />

tissue recombinants similar to those described to demonstrate<br />

that both stomal and epithelial ER� are required for<br />

estrogen induction of the uterine epithelial secretory prod-<br />

ucts, lactoferrin and complement component C3. Therefore,<br />

estrogen-induced proliferation of the uterine epithelium requires<br />

the presence of ER� in the stromal compartment only,<br />

whereas induction of certain epithelial secretory products is<br />

dependent on the presence of ER� in both uterine compartments.<br />

3. Maintenance of selective estrogen actions in the �ERKO uterus.<br />

A distinct advantage of null receptor models, whether naturally<br />

existing or experimentally generated via molecular<br />

methodologies, is their use as an in vivo tool for discerning<br />

alternate pathways of hormone action. Recent studies by the<br />

Lubahn laboratory have indicated the preservation of a distinct<br />

estrogen-signaling pathway in the �ERKO uterus (173,<br />

174). Das et al. (173) reported that two consecutive treatments<br />

(over a period of 12 h) with the catecholestrogen, 4-hydroxyestradiol<br />

(4-OH-E 2)at10�g/kg body weight resulted<br />

in significant increases in water imbibition in the uteri of<br />

ovariectomized �ERKO mice. Induction of lactoferrin<br />

mRNA in the uterine epithelium of wild-type mice was 97and<br />

85-fold after treatment with 10 �g/kg body weight estradiol<br />

or 4-OH-E 2, respectively (173). In contrast, only the<br />

4-OH-E 2 was active in the �ERKO uterus, resulting in a<br />

60-fold increase in lactoferrin mRNA levels compared with<br />

a 1.4-fold induction by estradiol (173). A similar, yet more<br />

modest, response was reported in the wild-type and �ERKO<br />

uteri after treatment with the xenoestrogens, kepone (15<br />

mg/kg body weight) (173) and methoxychlor (15 mg/kg<br />

body weight) (174).<br />

Most interesting was the lack of inhibition of this response<br />

by the pure estrogen antagonist, ICI-182,780, in both the<br />

wild-type and �ERKO mice, indicating the possibility of a<br />

non-ER�-mediated signaling pathway for certain compounds<br />

exhibiting estrogenic activity. Additionally, the nature<br />

of the timing and type of lactoferrin response elicited by<br />

the 4-OH-E 2 and xenoestrogens in the �ERKO is quite distinct<br />

from that of the original descriptions by Teng et al. (175)<br />

concerning estrogen regulation of this gene in the mouse<br />

uterus. Given the knowledge of the low-to-absent expression<br />

of ER� in the uterus and the ability of the ICI compounds to<br />

antagonize ER� signaling in vitro, it is not likely that this<br />

receptor is involved in this phenomenon. The catecholestrogens<br />

are naturally synthesized and proposed to play a role<br />

in steroid regulation of the hypothalamus and pituitary (176,<br />

177), ovarian function (178), and embryo implantation (179).<br />

Furthermore, the discovery of local synthesis of these estrogens<br />

in mammary tissue has led to implications of their<br />

involvement in breast cancer (180). Therefore, further investigation<br />

into the alternate mechanisms by which these compounds<br />

may activate nuclear processes is needed.<br />

4. Maintenance of progesterone action in the �ERKO uterus. Like<br />

estradiol, ovarian derived progesterone, is an integral steroid<br />

hormone in the physiology and function of the uterus. The<br />

PR has been localized to cells composing all three anatomical<br />

compartments of the uterus and exhibits varied levels in each<br />

during the stages of the estrous cycle (reviewed in Ref. 181).<br />

The PR also exists in two forms, PR A and PR B, which differ<br />

only in the length of the N� terminus. In contrast to the ER,<br />

PR A and PR B are encoded by a single gene but transcribed


372 COUSE AND KORACH Vol. 20, No. 3<br />

from two distinct promoters in both the rat (160) and human<br />

(159).<br />

As previously discussed, the PR gene is strongly regulated<br />

by the estradiol-ER� complex. This is evidenced by in vitro<br />

and in vivo studies showing inhibition of estrogen stimulation<br />

of the PR gene with antiestrogens as well as the presence<br />

of a single imperfect estrogen-response element in the proximal<br />

region of the PR gene promoter (160). However, in vitro<br />

assays indicate that both the proximal as well as a distal<br />

promoter appear to be at least partially ER� dependent and<br />

may also involve ligand-independent pathways of ER� action<br />

for full expression (160, 182). The lack of estradiol-induced<br />

increases in PR mRNA levels in the �ERKO uterus<br />

confirms a regulatory dependence of the PR gene on ER�<br />

action (123). Therefore, it was hypothesized that disruption<br />

of the ER� gene may subsequently result in abnormally low<br />

levels of PR in the �ERKO uterus, and thereby render this<br />

tissue refractory to progesterone as well. However, Northern<br />

blot and ribonuclease protection assays have indicated basal<br />

levels of PR mRNA in the �ERKO uterus that are equal to<br />

those in wild-type, although estrogen-stimulated increases in<br />

these levels are absent (123). Binding assays with a radiolabeled<br />

progestin indicate levels of PR protein in the �ERKO<br />

uteri are present but reduced to approximately 60% of that<br />

in wild type (183). A notable finding was the greater proportion<br />

of PR found to be tightly associated with the nuclei<br />

of cells from the �ERKO uteri (�25%), compared with the<br />

wild type (�5%) (183). It is possible that a lack of ER� has<br />

resulted in a more plentiful pool of available coregulator<br />

proteins, allowing the PR present in the �ERKO uterus to<br />

maintain a more “active” state. Western blots indicate no<br />

difference in the relative levels of PR A and PR B between the<br />

two genotypes, with PR A consistently present in greater<br />

amounts in both (183). Therefore, a loss of ER� action in the<br />

uterus has led neither to a complete loss of PR nor to altered<br />

and preferential transcription from one of the two PR gene<br />

promoters.<br />

Curtis et al. (183) carried out a series of studies demonstrating<br />

the preservation of PR-mediated progesterone actions<br />

in the �ERKO uteri. Stimulation of the genes encoding<br />

amphiregulin (184) and calcitonin (185) in the rodent uterus<br />

has been shown to be an early and late response to progesterone,<br />

respectively. A treatment regimen of progesterone<br />

shown to be effective in inducing these two genes in the<br />

ovariectomized wild-type uterus was equally effective in the<br />

uteri of ovariectomized �ERKO females (183). Furthermore,<br />

the documented ability of estradiol to inhibit the progesterone<br />

induction of the amphiregulin gene was absent in the<br />

�ERKO (183). These studies indicate that a sufficient level of<br />

PR and an active progesterone signaling pathway are present<br />

in the uteri of �ERKO mice.<br />

A well known physiological role for progesterone is the<br />

preparation of the uterine endometrium for the forthcoming<br />

pregnancy. Implantation of the embryo is a complex process<br />

that requires a high degree of synchronicity between the<br />

blastocyst and the hormone-induced changes of the uterine<br />

endometrium (reviewed in Ref. 186). In general, ovarian<br />

synthesis of estradiol, which peaks at ovulation, serves to<br />

“prime” the uterus by eliciting increased differentiation and<br />

proliferation of the luminal and glandular epithelium, and<br />

the induction of significant increases in PR in the endometrial<br />

stroma and myometrium (162, 186). Subsequent increases in<br />

progesterone released from the ovarian corpora lutea then<br />

cause decidualization, a complex process involving massive<br />

proliferation and differentiation of the endometrial stroma<br />

along with localized increases in vascular permeability and<br />

edema (186). This process involves the synthesis of specific<br />

hormones, such as PRL, cytokines, prostaglandins, extracellular<br />

matrix components, and various enzymes within the<br />

uterine tissues (186). The result is a remarkable swelling of<br />

the uterine stroma thought to be necessary for implantation<br />

by forcing the uterus to close down on the blastocyst (186).<br />

The final stage of apposition in the mouse is the grasping of<br />

the blastocyst and ultimate attachment to the uterine wall, a<br />

process thought to be dependent on the secondary rises in<br />

ovarian derived estradiol (186). Therefore, preparation of the<br />

uterus for blastocyst implantation is dependent on the multifunctional<br />

and sometimes opposing effects of estradiol and<br />

progesterone.<br />

A laboratory model for the decidualization process has<br />

been generated for several species and generally involves<br />

exogenous steroidal treatments designed to mimic those of<br />

the ovarian cycle, coupled with a mechanical stimulus or<br />

trauma of the uterine lining to simulate implantation (186,<br />

187). In brief, ovariectomized mice are treated with estradiol<br />

(100 ng) for 3 consecutive days, followed by 2 days of no<br />

treatment. The animals are then administered progesterone<br />

(1 mg) and a reduced dose of estradiol (7 ng) for 8 consecutive<br />

days, with mechanical stimulation of the uterus taking place<br />

on the third day. Using this treatment scheme and the progesterone<br />

receptor-knockout (PRKO) model, Lydon et al. (44)<br />

demonstrated the absolute dependence of the decidualization<br />

process on progesterone action and the PR by reporting<br />

the complete lack of decidualization in the PRKO mice. However,<br />

Curtis et al. (183) demonstrated that a progesteroneinduced<br />

decidual response is possible in the �ERKO uterus,<br />

despite its inability to respond to the estradiol priming meant<br />

to mimic estrus. Most interesting were investigations indicating<br />

that although the uterine decidualization in the wild<br />

type was dependent on estradiol, as evidenced by its inhibition<br />

with ICI-182,780, progesterone alone was sufficient to<br />

induce decidualization in the �ERKO uterus (183).<br />

The reasons for the apparent loss of estrogen dependence<br />

for successful decidualization in the �ERKO uterus can only<br />

be speculated upon at this time. One role of estrogen in<br />

uterine decidualization is thought to be the induction of<br />

significant increases in PR in the endometrial stroma (186).<br />

This is supported by recent immunohistochemical studies by<br />

Tibbetts et al. (162) demonstrating strong up-regulation of<br />

stromal PR and simultaneous down-regulation of epithelial<br />

PR in the mouse uterus after 4 days of estradiol treatment.<br />

In light of the puzzling results indicating the PR in the<br />

�ERKO uterus is strongly associated with the nucleus, and<br />

therefore possibly in a more “active” state, the ER�-independent<br />

decidualization observed in the �ERKO uterus may<br />

be due to an enhanced ability to respond to progesterone.<br />

However, a release of suprabasal levels of estradiol during<br />

the luteal phase of the ovarian cycle is synchronized with the<br />

time of implantation and thought to be critical to enhancing<br />

and maximizing the uterine decidualization process (186).


June, 1999 ESTROGEN RECEPTOR NULL MICE 373<br />

Nonetheless, the deciduomas induced in the �ERKO females<br />

are neither reduced in size nor appear less complex or differentiated<br />

than those observed in the wild type (183). It is<br />

also possible that a lack of ER� during development as well<br />

as adulthood has resulted in a uterus with a heightened<br />

tendency toward decidualization, caused by to the altered<br />

expression of other gene products. The complexity of the<br />

decidual process is illustrated by the several models that lack<br />

the ability to exhibit uterine decidualization, such as mice<br />

lacking leukemia-inhibitory factor (188), prostaglandin synthase-2<br />

(189), and Hoxa-10 (190). Furthermore, a process<br />

thought to be critical to implantation is the acquired ability<br />

of portions of the uterine epithelium to self-destruct and<br />

become detached from the uterine wall, possibly clearing a<br />

route by which the underlying swelling endometrium can<br />

breach and provide a site for implantation (186, 191). Histological<br />

analysis of �ERKO uteri indicates a uterine epithelium<br />

that may be less healthy and more often exhibits sloughing<br />

compared with wild-type uteri. It is therefore possible<br />

that the inherently impaired luminal epithelium of the<br />

�ERKO female has resulted in a lowering of the threshold<br />

required to induce decidualization.<br />

B. Vagina<br />

The fully developed adult vagina serves as both a copulatory<br />

receptacle and a birth canal in the female and may be<br />

divided into two distinct sections, the upper vagina and<br />

lower vagina. The cranial end of the upper vagina is attached<br />

to the cervix and is derived from the Müllerian ducts during<br />

differentiation of the female tract (143). The lower vagina,<br />

which connects the tract to the vulva and external genitalia,<br />

is differentiated from the urogenital sinus (143). As shown in<br />

Fig. 3, the wild-type vagina is a highly sensitive estrogen<br />

target tissue, composed of an inner mucosal layer of stroma<br />

and overlying epithelia, a middle layer of muscularis, and an<br />

outer sheath of connective tissue. Detectable levels of ER� are<br />

present in both the stromal and epithelial compartments<br />

making up the mucosa of the duct (7). Estradiol exposure<br />

during the ovarian cycle induces a series of effects in the<br />

vaginal mucosa that are often used to estimate serum gonadal<br />

hormone levels and approximate the current stage in<br />

the estrous cycle (107). These changes in the mucosa include<br />

cytodifferentiation of the stromal cells and a rapid proliferation<br />

and differentiation of the epithelial cells, resulting in a<br />

stratified and cornified epithelial layer closest to the lumen<br />

(192). This process also involves the estrogen stimulation of<br />

a series of keratins (193, 194). As shown in Fig. 3, despite the<br />

chronically elevated levels of estradiol in the serum of<br />

�ERKO females, histological analysis consistently indicates<br />

a complete lack of vaginal estrogenization. A similar effect on<br />

the vaginal mucosa has been produced in mice (195, 196) and<br />

rats (197) after prolonged ovariectomy or treatments with the<br />

antiestrogens, ZM-189,154, EM-800, and tamoxifen. Administration<br />

of exogenous estradiol, DES, or hydroxytamoxifen<br />

to �ERKO mice results in no discernible vaginal response<br />

(153). In contrast, the vaginal mucosa of the �ERKO female<br />

appears to undergo the normal cyclic changes associated<br />

with ovarian steroidogenesis (Fig. 3) (47), strongly indicating<br />

that this is an ER�-mediated process.<br />

Buchanan et al. (198) have used the stromal-epithelial tissue<br />

recombinant scheme described above for uterine tissue to dissect<br />

the contributions of the different tissue compartments in<br />

the estradiol response of vaginal tissue as well. As in the uterus,<br />

these studies demonstrated that stromal ER�, but not epithelial<br />

ER�, are required for estradiol-induced epithelial proliferation<br />

in the mouse vagina (198). Similar to the observations in the<br />

uterine recombinants, both vaginal stromal and epithelial ER�<br />

were required for estradiol-induced stratification and cornification<br />

of the epithelium, including the induction of the gene for<br />

the secretory protein cytokeratin 10 (198). All recombinations<br />

involving tissue from the �ERKO vagina became atrophied,<br />

even in the presence of estradiol (198).<br />

C. Oviduct<br />

The mouse oviduct is a coiled tubular organ connecting the<br />

uterus to the ovarian bursa and derived from the Müllerian<br />

duct during fetal development of the female reproductive<br />

tract (143). It functions as a route for passage of sperm to the<br />

ovulated oocyte and the subsequently fertilized blastocyst to<br />

the uterus. In the CD-1 mouse, ER� immunoreactivity is<br />

easily detectable in both the stroma and the epithelium of the<br />

oviduct as early as day 2 of life (144). The levels of ER�<br />

immunoreactivity in the epithelium continue to rise and<br />

plateau at approximately day 15 and remain high throughout<br />

adulthood (144). Furthermore, Newbold et al. (199, 200) described<br />

the high degree of sensitivity of the fetal and neonatal<br />

oviduct to the detrimental effects of developmental DES<br />

exposure. During adulthood, the levels of ER� in the oviductal<br />

epithelium fluctuate during the ovarian cycle, reaching<br />

peak levels during the proliferative phase (201). The<br />

ovarian sex hormones found in high concentrations in the<br />

oviductal fluid are thought to play a role in ovum and zygote<br />

transport through the oviduct (201). However, studies using<br />

ovariectomized laboratory animals have produced conflicting<br />

results in terms of what this role may be, depending on<br />

the animal model and hormone dosing regimen used (202).<br />

Despite the apparent ontogeny of ER� in the developing and<br />

adult mouse oviduct, no gross phenotypes in the oviduct of<br />

�ERKO females have been observed. Similar to the uterus,<br />

the epithelium of the �ERKO oviduct usually appears<br />

healthy yet unstimulated, despite the chronically high levels<br />

of serum estradiol. Due to the inability of the �ERKO to<br />

ovulate, possible defects in the transport functions of the<br />

oviduct are not easily studied. Assays for ER� mRNA in the<br />

mouse oviduct detect little if any ER� transcripts. Accordingly,<br />

in �ERKO females there appears to be no obvious<br />

defects in the structure and function of the oviduct that<br />

impede fertility.<br />

D. Ovary<br />

In most mammals, differentiation of the bipotential fetal<br />

gonad to an ovary in the genotypic female occurs later in<br />

gestation than differentiation of the testis in the male (111).<br />

The factors involved in development and differentiation of<br />

the ovary are not well understood, although the process does<br />

not appear to be dependent on the presence of primordial<br />

germ cells (111). The appearance of follicles and the onset of


374 COUSE AND KORACH Vol. 20, No. 3<br />

estrogen synthesis in the fetal gonad may be the first indication<br />

of differentiation to an ovary, although the secreted<br />

estrogens do not appear to be critical to development of the<br />

ductal structures of the female reproductive tract. Still, fetal<br />

ovarian estradiol may play a role in development of the<br />

ovary itself, as evidenced by the complete lack of ovaries in<br />

SF-1 knockout mice (137, 138). Furthermore, a recent study<br />

has demonstrated immunohistochemical detection of ER�<br />

and ER� in the neonatal rat ovary (103). Interestingly, a lack<br />

of ER� or ER� during development appears to have no gross<br />

effect on ovarian differentiation, since individual knockouts<br />

of both respective receptors possess normal ovaries at birth<br />

and during prenatal development (47, 142). A study of ovarian<br />

development in a double-knockout lacking both forms of<br />

ER will therefore prove interesting in the future. Still, distinct<br />

ovarian phenotypes become apparent in the adult �ERKO<br />

and �ERKO females, resulting in infertility and subfertility<br />

in each, respectively (46, 47).<br />

1. Review of the physiology and function of the ovary. A brief<br />

description of ovarian morphology is necessary for a discussion<br />

of phenotypes that result from a lack of ER. The<br />

ovary may be conveniently divided into three broad functional<br />

units: the follicles, corpora lutea, and interstitial/stromal<br />

compartment (203, 204). All three possess the capacity to<br />

synthesize hormonal factors, especially steroids, in response<br />

to gonadotropins secreted from the anterior pituitary. The<br />

maturing follicle is a relatively ellipsoidal unit that can be<br />

further divided into three main cell types: the outermost<br />

thecal cells, which surround a single or multiple layers of<br />

granulosa cells, and together act to encase the germ cell<br />

(oocyte) at the approximate core. The overall size of the<br />

follicle and the number of cells composing the thecal and<br />

granulosa cell compartments are dependent on the stage of<br />

maturation (204). The corpora lutea are clearly defined and<br />

vascularized structures formed from the terminally differentiated<br />

thecal and granulosa cells that remain after ovulation<br />

(203). And finally, the interstitial and stromal tissue is<br />

composed of undifferentiated cells that may eventually be<br />

recruited for the thecal or granulosa units as well as dedifferentiated<br />

thecal and granulosa cells from past atretic follicles<br />

or regressed corpora lutea. This compartment also<br />

functions as the matrix within which the follicles are suspended<br />

(203).<br />

Ovarian function is often divided into two separate<br />

phases. The follicular phase refers to the period of follicle<br />

maturation and increased estradiol synthesis that leads up to<br />

and terminates with ovulation of the ovum. Ovulation marks<br />

the beginning of the luteal stage in which the developing<br />

corpora lutea secrete large amounts of progesterone as well<br />

as estradiol to allow for successful implantation of the blastocysts<br />

in the uterus. During the follicular stage of the ovarian<br />

cycle, the follicles may be categorized based on size,<br />

responsiveness to gonadotropins, and steroidogenic capabilities.<br />

These stages are most commonly referred to as the<br />

primordial, primary, secondary, tertiary or antral, atretic,<br />

and mature Graafian follicle (204). The primordial, or nongrowing<br />

follicles, are the most prevalent stage in the ovary<br />

at any one time and provide the pool from which follicles will<br />

be selected for maturation. These follicles consist of an oocyte<br />

arrested at the diplotene stage of the first meiotic division,<br />

surrounded by a single layer of cuboidal granulosa cells<br />

(204). Commencement of the follicular phase of the ovarian<br />

cycle involves the recruitment of primordial follicles to form<br />

the assembly of primary growing follicles to be prepared for<br />

ultimate ovulation. The factors required for this recruitment<br />

are not well understood. Henceforth, each stage is characterized<br />

by dramatic changes in the structure and functional<br />

capabilities of the follicle, which have been well characterized<br />

in several reviews (204–208). As the follicle progresses<br />

toward the secondary stage, rapid proliferation of the granulosa<br />

cells results in the formation of several concentric layers<br />

surrounding the oocyte (204). By this stage, stromal cells<br />

have differentiated to produce a defined stratum of thecal<br />

cells that encapsulate the granulosa cell/oocyte unit. A basement<br />

membrane acts to separate the heavily vascularized<br />

thecal layers from the granulosa cells and ovum. Since capillaries<br />

do not penetrate the basement membrane, the granulosa/oocyte<br />

compartment depends on the passive movement<br />

of hormonal factors through this extracellular structure<br />

(204). Oocyte and follicular growth are linear up to the tertiary<br />

stage, at which time the ooctye appears to reach a<br />

maximum size, while the follicle as a unit continues to enlarge.<br />

The tertiary follicle is characterized by a hypertrophied<br />

thecal layer, multiple layers of granulosa cells, and the appearance<br />

of an antrum, a space that separates the granulosa<br />

cells from the ooctye/cumulus complex. The process of follicular<br />

selection for ovulation, although not well understood,<br />

appears to occur at this stage of folliculogenesis, when several<br />

secondary-tertiary follicles will divert toward a pathway<br />

of atresia. Still under the influence of gonadotropins, the<br />

“selected” follicles will continue to enlarge, mostly due to<br />

increases in antrum size, to eventually reach the Graafian, or<br />

ovulatory stage. Follicular rupture and ovulation occur in<br />

response to a surge in gonadotropin levels, at which time<br />

cellular proliferation is ceased, and the remaining cells of the<br />

follicle terminally differentiate to form the corpus luteum<br />

(209).<br />

In consonance with its gametogenic function, the ovary<br />

fulfills a critical role as an endocrine organ, serving as the<br />

principal source of sex steroids in the female. Therefore, a<br />

normal functioning ovary is an essential prerequisite to the<br />

function and maintenance of the reproductive tract, mammary<br />

gland, and behavior of the female. The research efforts<br />

of several investigators during the past decades have generated<br />

the well accepted “two-cell, two-gonadotropin”<br />

model of ovarian estradiol synthesis. This model and the<br />

investigations leading to its description have been discussed<br />

in great detail in several recent reviews and therefore will be<br />

only summarized here (204, 206, 210). The two steroid-producing<br />

components of the maturing follicle are the thecal and<br />

granulosa cells, which predominantly produce androgens<br />

and estrogens, respectively (210). Ample evidence exists to<br />

indicate that thecal cells possess the full complement of steroidogenic<br />

enzymes necessary for estradiol synthesis. In contrast,<br />

estradiol synthesis by the granulosa cells is dependent<br />

on thecal-derived androgens as substrates for aromatization.<br />

The amount and activity of the expressed steroidogenic enzymes<br />

within the two cell types vary depending on the<br />

follicular stage, and thereby determine the predominant ste-


June, 1999 ESTROGEN RECEPTOR NULL MICE 375<br />

roid being produced. The cell-specific and temporal actions<br />

of the gonadotropins, LH and FSH, regulate the type and<br />

activity of the steroidogenic enzymes expressed within the<br />

granulosa and the thecal cells. The model states that LH<br />

acting via the constituitively expressed LH receptor on the<br />

cell surface of thecal cells stimulates the synthesis of androgens<br />

(androstenedione) in the growing follicle. This requires<br />

the initial conversion of cholesterol stores to pregnenolone by<br />

the cholesterol side-chain cleavage enzyme (P450 scc) and is<br />

thought to be a rate-limiting step in thecal cell steroidogenesis<br />

(204). Still within the thecum, pregnenolone is converted<br />

to progesterone and then to androstenedione via the enzymatic<br />

actions of 3�-hydroxysteroid dehydrogenase and 17�hydroxylase/C<br />

17–20 lyase (P450 17�), respectively (204, 206,<br />

210). Regulation by LH has been shown to occur at both the<br />

transcriptional and translational levels for the P450 scc and<br />

P450 17� genes (206). Granulosa cells lack expression of the<br />

P450 17� enzymes required to produce androgens, the precursor<br />

of estradiol, and therefore are dependent on the passage<br />

of the thecal-derived androgens through the basement<br />

membrane and into the granulosa compartment. This cellular<br />

cooperation provides the basis of the two-cell portion of<br />

the model. The second gonadotropin, FSH, acts solely upon<br />

the granulosa cells to stimulate the enzymatic conversion of<br />

the androstenedione and testosterone to estrone and estradiol,<br />

via P450-aromatase (P450 arom), and 17�-hydroxysteroid<br />

dehydrogenase, respectively (204, 206, 210). The estradiol is<br />

then released into the follicular fluid, whereupon the bulk<br />

passes back through the basement membrane and enters the<br />

circulation. Upon ovulation, the luteal phase begins with<br />

luteinization of the follicle and differentiation of the remaining<br />

granulosa and thecal cells to form the corpus luteum. The<br />

relative amounts and activities of the steroidogenic enzymes<br />

are altered once again and shift toward synthesis of large<br />

amounts of progesterone.<br />

Recent data have challenged the “two-cell” model to<br />

incorporate the descriptions of a role for the oocyte in regulating<br />

granulosa cell steroidogenesis. Elegant in vitro experiments<br />

involving the surgical removal of the oocyte from<br />

isolated growing follicles have demonstrated the existence of<br />

an ooctye-secreted factor that is able to inhibit granulosa cell<br />

estradiol and progesterone synthesis (211–213).<br />

2. Review of intraovarian estrogen actions. In 1940, both Pencharz<br />

(214) and Williams (215) independently reported a<br />

direct and specific ability of estradiol or DES to induce significant<br />

increases in ovarian weight in the hypophysectomized<br />

rat. These same seminal studies also described the<br />

synergistic effect of estradiol on gonadotropin-stimulated<br />

increases in ovarian weight (214, 215). Since then, numerous<br />

intraovarian effects of large amounts of locally synthesized<br />

estrogens have been described and postulated to be essential<br />

to normal follicular development and ovarian function. In<br />

granulosa cells of the growing follicle, estrogen has been<br />

reported to increase the levels of its own receptor (216), as<br />

well as induce DNA synthesis and proliferation (205, 217–<br />

220), increase the number and size of intercellular gap junctions<br />

(221), stimulate synthesis of IGF-I (222), and attenuate<br />

apoptosis and follicular atresia (223, 224). Estradiol is also<br />

known to augment the actions of FSH on granulosa cells,<br />

resulting in the maintenance of FSH-receptor levels (218,<br />

225–227) and the acquisition of LH-receptor (218, 228–231),<br />

an event critical to successful ovulation.<br />

Ultimately, the actions of estradiol act to enhance follicular<br />

responsiveness to gonadotropins and thereby result in increased<br />

aromatase activity and further estrogen synthesis<br />

(231, 232). Therefore, normal ovarian function appears to be<br />

dependent on a multitude of auto- and paracrine actions of<br />

estradiol that act in concert with the gonadotropins secreted<br />

from the anterior pituitary to provide for successful folliculogenesis<br />

and steroid production. Nonetheless, immunohistochemical<br />

detection and characterization of ER in the different<br />

ovarian compartments have proven difficult, although<br />

studies employing binding assays with radiolabeled ligands<br />

report the presence of ER in ovarian granulosa cells of the rat<br />

(216, 233–235), mouse (236), rabbit (236), and pig (235).<br />

The discovery of the ER� and its reportedly high mRNA<br />

levels in the ovary (49, 63, 93) reinforces the need for thorough<br />

immunohistochemical studies for the two distinct ERs<br />

in the ovary. Reports of localization of ER� and ER� transcripts<br />

in the rat ovary by in situ hybridization indicate the<br />

presence of low levels of ER� mRNA with no specific pattern<br />

(63), whereas ER� mRNA is easily detectable and predominantly<br />

localized to the granulosa cells of growing follicles<br />

(49, 63). Sar and Welsch (103) recently described immunohistochemical<br />

studies with ER�- and ER�-specific antibodies,<br />

reporting that ER� immunoreactivity is indeed highly<br />

expressed in and localized to the granulosa cells of growing<br />

follicles, whereas ER� staining appears limited to the interstitial/thecal<br />

cells in the rat ovary. Similar findings of immunohistochemical<br />

localization of the ER� to the ovarian<br />

granulosa cells were reported in the rat by Hiroi et al. (104)<br />

and in the cow by Rosenfeld et al. (69). Brandenberger et al.<br />

(94) reported the RT-PCR detection of ER� and ER� transcripts<br />

in normal and neoplastic human ovary and ovarian<br />

cell lines. This study further described the presence of easily<br />

detectable levels of ER� mRNA and very low levels of ER�<br />

mRNA in the granulosa cells, whereas the opposite was<br />

found in a cell-line derived from the ovarian outer surface<br />

epithelium (94). Misao et al. (237) also used RT-PCR to detect<br />

ER� and ER� transcripts in human corpus luteum. Both<br />

Misao et al. (237) and Byers et al. (63) demonstrated a downregulation<br />

of ER� mRNA during luteinization of the follicle<br />

and the differentiation of the corpus luteum in the human<br />

and rat, respectively. Interestingly, a report by Iwai et al.,<br />

before the knowledge of ER�, described the detection of ER�<br />

immunoreactivity in the granulosa cells of the rabbit ovary<br />

(238), possibly illustrating another variation in the expression<br />

pattern of the two ERs among different species. Nonetheless,<br />

ER� and ER� are present in the adult rodent ovary.<br />

Therefore, disruption of the genes encoding these receptors<br />

may be expected to result in distinct ovarian phenotypes. In<br />

addition, the dissimilar expression patterns for ER� and ER�<br />

among the functional units of the follicle suggest that compensatory<br />

mechanisms fulfilled by the remaining functional<br />

gene in each respective ERKO may not be possible in the<br />

ovary.<br />

3. �ERKO ovary. The ovary of the neonatal and prepubertal<br />

�ERKO female does not exhibit any gross differences when


376 COUSE AND KORACH Vol. 20, No. 3<br />

compared with those of wild-type littermates (142). The mature<br />

�ERKO ovary possesses a normal complement of primordial<br />

follicles, indicating no defects in germ cell generation<br />

or migration to the gonad during fetal development.<br />

However, at the commencement of sexual maturity, it becomes<br />

apparent that the �ERKO female is anovulatory and<br />

exhibits a distinct ovarian phenotype of enlarged, hemorrhagic,<br />

and cystic follicles as shown in Fig. 4. These cystic<br />

structures often accumulate in the ovary, making the gonad<br />

appear grossly as a bundle of dark grapes, a signature phenotype<br />

of the �ERKO female. However, growing follicles in<br />

the tertiary and pre- to small antral stage are present in the<br />

mature �ERKO ovary, indicating that ER� is not required for<br />

the recruitment of primordial follicles and the initial stages<br />

of the follicular phase. This is in contrast to the described<br />

phenotypes of mice possessing a mutation of the steel (Sl/Sl t )<br />

locus (239), or a targeted disruption of the genes encoding<br />

growth differentiation factor-9 (240), or the vitamin D receptor<br />

(45), all of which exhibit phenotypes of follicular arrest<br />

at very early stages. The growing and cystic follicles in the<br />

�ERKO are often characterized by a hypertrophied thecum,<br />

indicating stimulation by LH. Induction of thecal cell function<br />

by the chronically elevated serum LH is illustrated by the<br />

similarly elevated levels of serum androgens exhibited in the<br />

adult �ERKO female (Table 2). Histological analysis of ovaries<br />

from numerous �ERKO females has revealed no corpora<br />

lutea, indicating an inability of the follicles to spontaneously<br />

ovulate and differentiate. Anovulation in the adult �ERKO<br />

could not be rescued even after appropriate stimulation with<br />

exogenous gonadotropins (142), strongly indicating that infertility<br />

in the �ERKO female is due in large degree to alterations<br />

in the hormonal milieu and ovarian responsiveness.<br />

A phenotype of anovulation and follicular arrest is also<br />

observed in the ovaries of other murine models of gene<br />

disruption. The ovaries of mice lacking the actions of FSH,<br />

either due to targeted disruption of the gene for the FSH�subunit<br />

(241) or the FSH-receptor (242), exhibit follicles that<br />

grow up to but not beyond the preantral stage. Therefore,<br />

FSH action is critical to completion of the follicular phase of<br />

the ovarian cycle. However, �ERKO female mice possess<br />

serum FSH levels within the normal range (Table 2) and<br />

actually show elevated levels of FSH-receptor mRNA in the<br />

ovary (142). Furthermore, since FSH signaling has been proposed<br />

to be a critical factor in estradiol synthesis (243, 244),<br />

the extremely elevated levels of serum estradiol in the<br />

�ERKO female indicate that this pathway is intact in the<br />

granulosa cells. Interestingly, neither report of mice lacking<br />

FSH action mention any significant differences in ovarian<br />

estradiol synthesis but describe a much smaller and thinner<br />

uterus in the female (241, 242). It is possible that the estradiol<br />

present in these models lacking FSH action is synthesized by<br />

the ovarian thecum. Alternatively, the presence of compensatory<br />

mechanisms that have allowed for granulosa cell aromatase<br />

activity that is independent of FSH stimulation must<br />

also be considered. Progression of the follicle to the more<br />

mature antral stage is also arrested in mice after targeted<br />

disruption of the gene for IGF-I (245), cyclin-D2 (246), connexin-37<br />

(247), activin type II receptor (248), and superoxide<br />

dismutase 1 (249), although low serum gonadotropin levels<br />

are thought to be the cause of this phenotype in the latter two<br />

models.<br />

Therefore, a definitive cause of the �ERKO ovarian phenotype,<br />

in addition to the loss of ER� action, was not obvious.<br />

In fact, some facets of ovarian physiology thought to be<br />

dependent on estrogen action, such as the attenuation of<br />

apoptosis and the induction of LH receptors in granulosa<br />

FIG. 4. Histology of a representative adult wild-type and �ERKO ovary. Shown are ovarian cross-sections from wild-type (a) and �ERKO (b)<br />

females at low magnification (13.2�). Note the presence of follicles at both the follicular and luteal (corpora lutea; CL) stages of the ovarian<br />

cycle in the wild-type ovary, whereas the �ERKO ovary is characterized by the presence of large, hemorrhagic, and cystic follicles, a sparse<br />

number of follicles at the early stages of proliferation, and a lack of corpora lutea. Panels c–d illustrate �ERKO ovarian tissue at high power<br />

magnification (132�): c, a healthy secondary follicle showing a normal oocyte and nucleolus; d, interface of two cystic follicles, demonstrating<br />

the heterogeneity in the number of granulosa cells that line the cysts, e.g., the left cyst has a single layer, whereas the right cyst has several<br />

layers of granulosa cells (indicated by arrows); e, hypertrophied thecal cells lining a hemorrhagic cyst in an �ERKO ovary. Scale bar � 0.5 �m.


June, 1999 ESTROGEN RECEPTOR NULL MICE 377<br />

cells of antral follicles, were apparently preserved in the<br />

follicles of the �ERKO ovary. Follicular atresia in the ovary<br />

is a hormonally controlled process that is critical to oocyte<br />

selection. Although the factors that trigger atresia are not<br />

well understood, it is characterized by apoptosis of the granulosa<br />

cells of the follicle (reviewed in Refs. 223 and 224).<br />

Estradiol has been shown to be one of the many factors<br />

reported to protect the follicle from becoming atretic (250).<br />

Furthermore, androgens reportedly accelerate the process,<br />

and it may be an altered estrogen/androgen synthesis ratio<br />

in the follicle that leads to atresia (250). However, despite<br />

elevated androgen production, the presence of androgen<br />

receptor (AR) mRNA, and a lack of ER� action in the mature<br />

�ERKO ovary, an inordinate amount of apoptosis is not<br />

observed in the follicles (142). Estradiol has also been shown<br />

to facilitate the FSH induction of LH receptors in the granulosa<br />

cells of the mature ovulatory follicle in both in vivo (225,<br />

231, 252) and in vitro experiments (229, 230). Nonetheless, the<br />

granulosa cells of the growing follicles, in addition to the<br />

enlarged cysts in the �ERKO ovary, possess significant levels<br />

of LH receptor mRNA when assayed by in situ hybridization<br />

(142). The most plausible explanation for these observations<br />

is that these estrogen actions are mediated by ER�, which has<br />

been shown to be expressed in a normal pattern in the granulosa<br />

cells of the �ERKO ovary, and concomitant with the<br />

expression of LH receptor (93, 142).<br />

Therefore, with data suggesting that disruption of the ER�<br />

gene did not result in an ovary completely refractory to estrogen,<br />

other aspects of ovarian physiology must be considered as<br />

possible factors in the etiology of the �ERKO ovarian phenotype.<br />

As previously discussed, ovarian function is tightly controlled<br />

by pituitary gonadotropins (see Section III.D.1). In turn,<br />

gonadotropin synthesis and secretion from the anterior pituitary<br />

are at least partially regulated by gonadal steroids acting<br />

via classical feedback mechanisms in the hypothalamic-pituitary<br />

axis (reviewed in Ref. 251). Indeed, disruption of the ER�<br />

gene has resulted in significant phenotypes in the hypothalamic-pituitary<br />

axis of the �ERKO female (see Section VI.A). Most<br />

notable is the increased and chronic secretion of LH in the<br />

�ERKO female that results in levels that are 4–7 times that<br />

found in wild-type females (Table 2) (252). As discussed above,<br />

synchronized increases in serum FSH and LH levels are critical<br />

to follicular maturation and ultimate ovulation in the ovary.<br />

However, it has been proposed that the follicular requirements<br />

for LH are finite, and the presence of abnormally high levels<br />

may force maturing follicles to either prematurely luteinize or<br />

become atretic (253).<br />

Therefore, the �ERKO ovarian phenotype is likely caused<br />

by the chronic exposure to abnormally high levels of LH.<br />

Support for this hypothesis can be drawn from a number of<br />

studies. Investigations involving prolonged treatment with<br />

antiestrogens over a period of at least 28 days have produced<br />

an ovarian phenotype in both mice (195, 196) and rats (197)<br />

that is similar to that of the �ERKO. Of course, interpretation<br />

of these studies is complicated by the ability of the antiestrogens<br />

to inhibit both ERs as well as estrogen action in both<br />

the ovary and hypothalamic-pituitary axis. However, these<br />

studies reported that the “�ERKO” phenotype of enlarged<br />

cystic follicles was produced only after prolonged treatments<br />

with those antiestrogens that possessed the ability to cross<br />

the blood-brain barrier and concurrently produce serum LH<br />

levels that were several fold higher than controls. Therefore,<br />

whereas the estrogen antagonists ZM-189,154 (197) and EM-<br />

800 (195, 196) produced chronically elevated LH levels and<br />

an ovarian phenotype strikingly similar to that of the<br />

�ERKO, tamoxifen did neither (195–197). More definitively,<br />

Risma et al. (254, 255) report that targeted transgenic overexpression<br />

of the LH�-subunit in the mouse that results in a<br />

15-fold increase in serum LH levels produces females that are<br />

anovulatory and exhibit an ovarian phenotype almost indistinguishable<br />

from that of the adult �ERKO female.<br />

Therefore, the similarity in the ovarian phenotypes described<br />

in the above studies, in which the models presumably<br />

possessed normal levels of ovarian ER�, combined with our<br />

observations in the �ERKO, strongly indicate that the ER� is<br />

not directly involved in the development of ovarian cysts due<br />

to hypergonadotropism. However, there are descriptions of<br />

at least two models in which serum LH is chronically elevated,<br />

yet do not manifest an ovarian phenotype similar to<br />

the �ERKO or those induced by antiestrogens or transgenics<br />

as described above. Female mice that are homozygous for a<br />

targeted disruption of the FSH�-subunit gene exhibit an<br />

approximate 5-fold increase in serum LH but do not show<br />

indications of enlarged cystic follicles in the ovary (241),<br />

indicating a role for FSH in this process as well. Bogovich<br />

(256) has provided supporting evidence by demonstrating<br />

that FSH is required along with prolonged exposure to LH<br />

(in the form of human CG) to induce follicular cysts in the<br />

rat. A likely role for FSH is the induction of LH receptor in<br />

the granulosa cells of the maturing follicles, thereby rendering<br />

the follicle responsive to the increased levels of LH.<br />

Another contrasting knockout model is that of the P450 arom<br />

gene (ArKO), in which the homozygous females possess<br />

significantly elevated serum LH and FSH levels but lack the<br />

capacity to synthesize estradiol (257). Although folliculogenesis<br />

is arrested at an antral stage in the ArKO ovary, no<br />

�ERKO-like cystic structures are reported (257). Therefore,<br />

assuming that a lack of estradiol synthesis would disrupt<br />

ligand-dependent activity of the ER� in the granulosa cells,<br />

the lack of ovarian cysts in the ArKO indicate an intraovarian<br />

role for estradiol in this phenotype. Therefore, although the<br />

�ERKO phenoytpe may be triggered by hyperstimulation of<br />

the follicles by LH, it is likely influenced by both FSH and<br />

ER� actions in the granulosa cells as well. Current studies<br />

utilizing prolonged treatment of �ERKO females with a<br />

GnRH antagonist to reduce serum gonadotropins are being<br />

carried out to further define the etiology of the ovarian<br />

phenotype (J. F. Couse, D. O. Bunch, J. Lindzey, D. W.<br />

Schomberg, and K. S. Korach, manuscript in preparation).<br />

Since the �ERKO ovarian phenotype develops and worsens<br />

only after sexual maturity, we have began studies to<br />

characterize ovarian function in the immature �ERKO female<br />

(J. F. Couse, D. O. Bunch, J. Lindzey, D. W. Schomberg,<br />

and K. S. Korach, manuscript in preparation). Although superovulation<br />

with exogenous gonadotropins was not successful<br />

in eliciting ovulation in the older �ERKO females,<br />

immature (�28 days) females do respond and produce viable<br />

oocytes that can be collected from the oviduct. However, the<br />

average number of oocytes collected from superovulated<br />

�ERKO females is significantly less than that yielded from


378 COUSE AND KORACH Vol. 20, No. 3<br />

age-matched superovulated wild-type and heterozygous females.<br />

Therefore, intraovarian ER� action does not appear to<br />

be essential to ovulation when stimulated with exogenous<br />

gonadotropins.<br />

4. �ERKO ovary. As discussed above, the �ERKO has provided<br />

a number of indications to suggest that intraovarian<br />

ER� action is not critical to ovarian function. Furthermore,<br />

several presumed functions of estradiol in ovarian granulosa<br />

cells appear to be preserved in the �ERKO, such as the<br />

attenuation of apoptosis and the induction of LH receptors.<br />

Based on the reported localization of ER� mRNA (49, 50, 63)<br />

and protein (69, 103, 104) to the granulosa cells of growing<br />

follicles, as well as the maintenance of a normal expression<br />

pattern for ER� in the �ERKO ovary (93, 142), it is likely that<br />

the newly discovered ER is the predominant mediator of<br />

estrogen action in the follicle. In the rat ovary, ER� is easily<br />

detectable by immunohistochemistry and exhibits an almost<br />

ubiquitous expression pattern within the granulosa cells of<br />

follicles ranging from the primary up to the ovulatory or<br />

Graafian stage (69, 103, 104). Byers et al. (63) demonstrated<br />

that ER� mRNA levels remain relatively constant during the<br />

follicular stage of the ovarian cycle but are rapidly decreased<br />

after the gonadotropin surge-induced luteinization of the<br />

granulosa cells. Interestingly, in 1975, Richards (216) reported<br />

a similar profile for high-affinity [ 3 H]-E 2 binding in<br />

the granulosa cells of the rat ovary, showing increased binding<br />

levels as follicles matured followed by marked decrease<br />

after luteinization. Because the affinities of the two ERs for<br />

estradiol do not significantly differ, it was not possible at that<br />

time to realize that the receptor being detected in the ovary<br />

was distinctly different (i.e., ER�) from that which was being<br />

concurrently described in the uterus (i.e., ER�).<br />

As stated previously, female mice homozygous for the<br />

targeted disruption of the ER� gene possess no gross aberrant<br />

phenotypes as neonates or during adulthood (47). However,<br />

during a continuous mating study of 8 weeks in which<br />

sexually mature wild-type or �ERKO females were housed<br />

with a known fertile wild-type male, a significant deficit in<br />

fertility in the �ERKO became obvious. As shown in Table<br />

3, �ERKO females produced substantially fewer litters as<br />

well as significantly less numbers of pups per litter when<br />

compared with their wild-type littermates (47). Whereas the<br />

average litter size among wild-type females was 8.8 (�2.5)<br />

pups per litter, this number was reduced to 3.1 (�1.8) in<br />

�ERKO females (Table 3) (47). Furthermore, two of the tested<br />

�ERKO females yielded no litters, despite the observation of<br />

seminal plugs on multiple occasions, suggesting that abnormal<br />

sexual behavior was not a cause for the infertility. Gross<br />

TABLE 3. Fertility and superovulation data in the �ERKO female mice<br />

Genotype<br />

analysis of the uteri from the �ERKO females used in this<br />

study demonstrated no indication of embryo resorption during<br />

gestation. Therefore, the observed subfertility in the<br />

�ERKO females did not appear to be due to uterine dysfunction<br />

that results in premature termination of pregnancy.<br />

However, a possible defect in embryo implantation due to a<br />

lack of ER� could not be assessed in these studies.<br />

The nature of the subfertility in the �ERKO female described<br />

above strongly suggested an ovarian phenotype. Gross analysis<br />

of ovaries from �ERKO females indicated no distinct differences<br />

in size or morphology when compared with those of<br />

wild-type females. As shown in Fig. 5, histological analysis of<br />

ovaries from sexually mature �ERKO females illustrated the<br />

presence of a relatively normal interstitial compartment and<br />

follicles at various stages of the follicular cycle, ranging from<br />

primordial and primary to those with a clearly defined antrum,<br />

all possessing the expected thecal shell. Therefore, as demonstrated<br />

for ER� by the �ERKO, ER� does not appear to be<br />

essential for the establishment of germ cell number or ovarian<br />

development. Several of the speculated intraovarian roles of<br />

estrogen were discussed previously and include a proposed<br />

critical role in proliferation of the granulosa cells in the maturing<br />

follicle (217–220, 231). However, the multiple large follicles<br />

present in the �ERKO ovaries indicate no marked differences<br />

in granulosa cell number. Furthermore, serum estradiol levels<br />

in adult �ERKO females do not appear to differ from agematched<br />

wild-type females, at 24.2 (�3.3) and 30.5 (�2.8) pg/<br />

ml, respectively. Biological evidence of estradiol synthesis in the<br />

�ERKO ovary is also illustrated by the apparently normal<br />

uterus and vagina, which exhibit the proper cyclic changes in<br />

morphology of a sexually mature wild-type mouse. Nonetheless,<br />

there were indications of an increased number of early<br />

atretic follicles and a sparse presence of corpora lutea in the<br />

�ERKO ovary when compared with the wild-type (47), suggesting<br />

that the observed subfertility in the �ERKO may be due<br />

to a reduction in completed folliculogenesis.<br />

A phenotype of compromised follicular maturation that<br />

more often terminates in atresia rather than ovulation, as<br />

described in the �ERKO female, is similar to that reported in<br />

a number of other mutant mice. As previously discussed,<br />

arrested folliculogenesis is described in knockout mice for<br />

other genes known to be expressed in granulosa cells, including<br />

the gene for FSH-receptor (242), P450 arom (ArKO)<br />

(257), IGF-I (245), and connexin-37 (247). It is therefore possible<br />

that a loss of ER� action has also resulted in alterations<br />

in the expression and/or function of one or more of these<br />

gene products. However, the normal serum estradiol levels<br />

and the appearance of estrogenic effects in the reproductive<br />

Continuous mating results Superovulation results<br />

n Litters per<br />

female (SEM)<br />

Pups per<br />

litter (SEM)<br />

n Oocytes per<br />

female (SEM)<br />

Wild-type 6 2.8 � 0.4 8.8 � 2.5 10 33.7 � 4.8 9–57<br />

Heterozygous nd nd nd 11 52.5 � 5.7 a<br />

20–77<br />

�ERKO 11 1.7 � 1.0 a<br />

3.1 � 1.8 b<br />

11 6.0 � 1.5 a<br />

nd, Not determined.<br />

0–13<br />

a<br />

P � 0.05, Student’s two tailed t-test vs. wild-type<br />

b<br />

P � 0.001, Student’s two tailed t-test vs. wild-type<br />

Range


June, 1999 ESTROGEN RECEPTOR NULL MICE 379<br />

FIG. 5. Histology of representative adult wild-type and �ERKO ovary. Shown are ovarian cross-sections from adult wild-type (a) and �ERKO<br />

(c) females and those from immature wild-type (b) and �ERKO (d) females after superovulation treatment at low power magnification (13.2�).<br />

Note the presence of follicles at various stages of the follicular phase in both the wild-type (a) and �ERKO (c) ovaries from the adult females,<br />

illustrating relatively little observable difference between the two genotypes. However, upon superovulation, a distinct phenotype becomes<br />

apparent in the �ERKO. Comparison of the superovulated wild-type (b) with the similarly treated �ERKO (d), indicates the presence of multiple<br />

corpora lutea (CL) in the wild-type whereas only two CL are obvious in the �ERKO. Most notable are the multiple numbers of unruptured<br />

ovulatory follicles present in the superovulated �ERKO ovary (indicated by arrows). High-power magnification (132�) of the adult �ERKO ovary<br />

illustrates follicles at progressive stages of maturation (e) and a healthy tertiary follicle showing an oocyte with nucleolus (f). High-power<br />

magnification (132�) of the superovulated immature �ERKO ovary illustrates a typical unruptured ovulatory follicle possessing several layers<br />

of granulosa cells and a centrally located oocyte (g), and a corpora lutea (h) indicating the successful luteinization and terminal differentiation<br />

of the follicle that occurs after ovulation. [Panels a–d reproduced with permission from Ref. 47.] Scale bar � 1 �m.<br />

tract of �ERKO females suggest that induction of the<br />

P450 arom gene by FSH in the granulosa cells appears intact.<br />

However, investigation of the expression of the above as well<br />

as other granulosa cell components is required to determine<br />

the cause of the ovulatory phenotype of the �ERKO female.<br />

To gain further insight into the ovarian phenotype of the<br />

�ERKO female, immature knockout and wild-type mice were<br />

stimulated to ovulate by administration of superphysiological<br />

levels of gonadotropins (PMSG and hCG) (47). After several<br />

independent trials, it became obvious that the ovulatory capacity<br />

of the �ERKO female was dramatically reduced, yielding<br />

an average of 6 (�1.5) oocytes per female vs. 33 (�4.8) and 52


380 COUSE AND KORACH Vol. 20, No. 3<br />

(�5.7) oocytes per female in the wild-type and heterozygous<br />

animals, respectively (Table 3) (47). Additionally, the cumulus<br />

mass that surrounded the ovulated follicles from the �ERKO<br />

females was consistently composed of a decreased number of<br />

cells and a lessened integrity when compared with ova yielded<br />

from wild-type controls. Most interesting was the histology of<br />

the ovaries from the superovulated �ERKO females, which<br />

indicated the presence of numerous preovulatory but unruptured<br />

follicles (Fig. 5). It therefore appeared that the follicles of<br />

the �ERKO ovary were able to respond to the proliferative<br />

effects of PMSG in terms of increased size and antrum formation.<br />

However, a severe deficit in the response to the gonadotropin<br />

surge (hCG), required to induce luteinization and rupture<br />

of the follicle, was obvious in the �ERKO. A small number<br />

of the selected follicles were able to be expelled, as evidenced<br />

by the presence of ova in the oviduct and of corpora lutea in the<br />

corresponding �ERKO ovary. Therefore, a lack of ER� resulted<br />

in a drastic reduction in ovulatory capacity, yet with incomplete<br />

penetrance. Until �ERKO females are treated and tested in a<br />

similar manner to the �ERKO studies described, it will be difficult<br />

to determine the precise role for each ER in the ovary.<br />

The observation of numerous unruptured Graafian follicles<br />

in the ovaries of superovulated �ERKO females is strikingly<br />

similar to the phenotype reported for mice possessing<br />

a targeted disruption of the cyclin-D2 gene (246). Cyclin-D2<br />

is a positive regulator of cell cycle progression that is highly<br />

expressed in granulosa cells (reviewed in Ref. 258). Robker<br />

and Richards (259) demonstrated strong up-regulation of the<br />

cyclin-D2 gene by estradiol in rat granulosa cells and suggested<br />

that this protein may be a downstream mediator of the<br />

synergistic actions of FSH and estradiol that result in increased<br />

granulosa cell numbers in the maturing follicle. The<br />

dramatic increases in cyclin-D2 induced by estradiol are evident<br />

in vivo at both the mRNA and protein levels in the rat.<br />

Furthermore, assays on primary granulosa cell cultures from<br />

the rat indicate that estradiol stimulation of the cyclin-D2<br />

gene can be inhibited by the estrogen antagonist, ICI-164,384,<br />

strongly suggesting that it is an ER-mediated process (259).<br />

Therefore, given that ER� is the predominant form of ER in<br />

the granulosa cells, disruption of the ER� gene may likely<br />

result in significant deficits in cyclin-D2 expression in the<br />

granulosa cells of the growing follicles in the �ERKO female.<br />

However, FSH is also able to stimulate increases in cyclin-D2,<br />

although the temporal pattern of regulation by FSH is distinct<br />

from that elicited by estradiol (259). Nonetheless, it is<br />

possible that FSH action in the follicles of the �ERKO ovary<br />

have provided for some degree of cyclin-D2 expression and<br />

thereby may explain the incomplete penetrance of the<br />

�ERKO ovarian phenotype.<br />

The �ERKO ovarian phenotype that becomes apparent<br />

after superovulation is also similar to that reported for<br />

knockout models of the genes for the PR (PRKO) (44) and<br />

prostaglandin synthase-2 (189). The dramatic increases in<br />

both PR (260) and prostaglandin synthase-2 (261) in the granulosa<br />

cells of the ovulatory follicle shortly after the gonadotropin<br />

surge have been well documented. Furthermore, the<br />

lack of follicular rupture in the respective knockout models<br />

supports a critical role for each of these components in ovulation<br />

(reviewed in Ref. 262). Lydon et al. (44) reported infertility<br />

in the PRKO and a consistent inability of super-<br />

physiological doses of hCG to induce follicular rupture.<br />

Although regulation of the PR gene is strongly influenced by<br />

estradiol in the uterus (see Section III.A), sufficient evidence<br />

exists to indicate that this may not be the case in the ovary.<br />

For example, although the wild-type ovary possesses extremely<br />

high intraovarian levels of estradiol and the presence<br />

of ER�, levels of PR mRNA and protein remain at a modest<br />

basal level in granulosa cells of the pre- and antral follicle<br />

(260). However, within 4–6 h after the gonadotropin surge,<br />

transcription of the PR gene has peaked at levels several fold<br />

that before the surge, only to return to near-basal levels<br />

within 20 h (260). The mechanism of this strong and transient<br />

induction of the PR gene by LH is known to include significant<br />

increases in intracellular cAMP, but may also involve<br />

phosphorylation of ER� and/or a coactivator, which then<br />

combine to act in a synergistic nature (262). Therefore, the<br />

elevation in PR levels in the granulosa cells of the ovulatory<br />

follicle that is critical to follicular rupture may be attenuated<br />

in the �ERKO ovary.<br />

Another possibility for a lack of spontaneous ovulation in<br />

the �ERKO female may be not intaovarian in nature, but<br />

rather due to altered gonadotropin synthesis and secretion<br />

from the hypothalamic-pituitary axis. The sex steroids play<br />

an important role as a positive regulator of the preovulatory<br />

surge (reviewed in Refs. 263 and 264). Although the exact<br />

mechanism of action by which estrogens may be involved is<br />

not well defined, studies have shown that estradiol can induce<br />

GnRH release from the hypothalamus as well as cause<br />

increases in the level of GnRH receptors in the anterior pituitary<br />

(263). The ER� may be the predominant form of ER<br />

in the pituitary of the adult female mouse (93); however, both<br />

ER� and ER� have been detected in various regions of the<br />

hypothalamus (88, 97, 265). Preliminary data in the �ERKO<br />

female indicate that tonic levels of serum LH are within the<br />

normal range. However, a lack of hypothalamic ER� may<br />

have reduced the potential for positive regulation by estradiol<br />

in the hypothalamic-pituitary axis and thereby may<br />

result in a reduction in the frequency and/or amplitude of<br />

the preovulatory gonadotropin surge. Nonetheless, the results<br />

of the superovulation studies described above, in which<br />

an artificial bolus of gonadotropin is administered to induce<br />

ovulation, indicate a severe phenotype that can be localized<br />

to the ovary of the �ERKO female.<br />

IV. Mammary Gland<br />

In mammals, the mammary gland is essentially undeveloped<br />

at birth and does not undergo full growth until the<br />

completion of puberty and, in fact, remains undifferentiated<br />

until pregnancy and lactation. Development of the mammary<br />

gland may be divided into five distinct stages: embryonic<br />

and fetal, prepubertal, pubertal, sexually mature adult,<br />

and pregnancy/lactation (reviewed in Ref. 266). The influential<br />

factors involved at each stage differ in both type and<br />

magnitude. Although the developmental factors involved<br />

during the embryonic and fetal stages of the female mammary<br />

gland are poorly understood, estrogen action does not<br />

appear to be essential (266). However, studies have shown<br />

that the fetal and neonatal mammary gland of rodents is


June, 1999 ESTROGEN RECEPTOR NULL MICE 381<br />

responsive to the gonadal steroids, although distinct strain<br />

differences are evident (reviewed in Ref. 267). In male rodents,<br />

all or portions of the fetal glandular structure are<br />

destroyed via the “masculinization” effects of testicular androgens,<br />

an effect that can be reproduced in males by prenatal<br />

exposure to testosterone (267). Aberrant exposure of the<br />

neonatal female mouse to estradiol, testosterone, 5�-dihydrotestosterone,<br />

or PRL has been reported to result in an<br />

apparent increased sensitivity of the gland to mammotrophic<br />

hormones during adulthood, leading to varied degrees of<br />

excessive ductal growth and differentiation (267, 269). The<br />

later four stages of mammary gland development occur after<br />

birth and terminate in a gland capable of milk production.<br />

These stages are strongly regulated by the endogenous ovarian<br />

steroid hormones and are characterized by massive<br />

growth of the glandular ducts that emanate from the nipple<br />

until they have progressed through the fat pad composing<br />

the bulk of the breast. Upon pregnancy and the onset of<br />

lactation, the gland undergoes dramatic differentiation to<br />

produce milk- secreting structures, termed alveoli, throughout<br />

the ductal network. Therefore, since the majority of mammary<br />

gland growth and differentiation occurs after birth, this<br />

tissue provides a unique tool within which the interactions<br />

of gonadal and peptide hormone systems may be studied.<br />

A. �ERKO phenotype<br />

The mammary gland of an adult wild-type female mouse<br />

consists of a network of epithelial ducts originating from the<br />

nipple and forming a tree-like structure. The growth of the<br />

epithelial ducts begins during prepubertal development and<br />

continues during puberty until the branches of the gland<br />

have reached the limits of the fat pad (266). This ductal<br />

elongation is via cap cell proliferation at the terminal end<br />

buds of the individual ducts, as they maintain close contact<br />

with the stromal fat pad through which they are progressing.<br />

Studies using ovariectomized models have indicated that<br />

estradiol and GH, locally mediated as IGF-I, are required for<br />

the development of this ductal structure (266, 268). The mammary<br />

glands of adult �ERKO female mice exhibit a phenotype<br />

similar to the glands of a newborn female, confirming<br />

the need for ER�-mediated estradiol actions for ductal<br />

growth (269). However, the �ERKO gland does possess the<br />

component structures necessary for mammary gland development,<br />

i.e., the epithelial and stromal portions, connective<br />

tissue, and a small rudimentary ductal tree. Therefore, embryonic<br />

and fetal development of the mammary gland in the<br />

mouse occurs independently of ER� actions. However, this<br />

may be strain dependent, since studies in the C57BL mice, the<br />

background strain of the �ERKO, have also reported little<br />

effect of neonatal steroid exposure on the mammary gland<br />

(267). Nonetheless, despite apparently unaltered GH levels<br />

and the existence of significantly elevated levels of serum<br />

estradiol, the mammary gland of the �ERKO female has lost<br />

the capacity to commence the pre- and postpubertal stages<br />

of growth.<br />

Estradiol has been shown to directly stimulate the formation<br />

of terminal end buds and stimulate cellular proliferation of the<br />

mammary ductal epithelium (270). Furthermore, this physiological<br />

effect can be inhibited by antiestrogens (271), indicating<br />

a receptor-mediated pathway of estrogen action. Although ER<br />

has been detected in both the ductal epithelial cells as well as<br />

the stromal tissue of the mammary gland (270, 272), the mechanism<br />

by which estrogens directly regulate growth remain unclear.<br />

The lack of ER in the outermost proliferating cap cells of<br />

the terminal end buds suggests that the effects of estradiol may<br />

be indirect and may involve the regulation of cell cycle genes<br />

and/or paracrine-acting peptide growth factors and their cognate<br />

receptors (270). Both EGF and transforming growth factor-�<br />

are thought to be critical to proper mammary growth and<br />

development and are at least partially regulated by estradiol via<br />

the ER (273–275). Support for a role of estradiol-induced growth<br />

factor activity is provided by the study of Xie et al. (276), in<br />

which transgenic mice overexpressing a dominant-negative<br />

form of the EGF-R exhibited an attenuated response to EGF<br />

action in the mammary gland in vivo. The virgin transgenic mice<br />

of this study exhibited severe deficits in ductal growth and<br />

branching in the mammary gland, which were overcome only<br />

with pregnancy, when ovarian steroids increased endogenous<br />

EGF expression several fold that found in virgin females (276).<br />

Further evidence comes from Ankrapp et al., who recently reported<br />

that estradiol-releasing pellets implanted into the mammary<br />

fat pad of ovariectomized mice result in significantly<br />

increased levels of EGF and PR and induce terminal end bud<br />

formation in the gland (277). However, implantation of EGFreleasing<br />

pellets induced similar increases in PR and terminal<br />

end bud formation, suggesting that EGF may mediate the mitogenic<br />

actions of estradiol (277). Therefore, this study demonstrated<br />

a scenario of ER�-EGF cross-talk similar to that previously<br />

described in the uterus (see Section III.A.2). In brief,<br />

treatment of mice with pellets releasing both EGF and the antiestrogen<br />

ICI-182,780 exhibited inhibition of the mitogenic effects<br />

of the growth factor; in turn, an anti-EGF antibody was<br />

able to neutralize the growth effects induced by an implanted<br />

estradiol-releasing pellet (277). Therefore, as was confirmed in<br />

the uterus through the use of the �ERKO mice, EGF is also able<br />

to mimic the mitogenic effects of estradiol in the mammary<br />

gland, but once again appears to require the presence of functional<br />

ER� to complete this mechanism.<br />

It is well known that the mammary fat pad serves not only<br />

as a matrix, but also provides biochemical signals and/or<br />

factors necessary for normal ductal growth (reviewed in Ref.<br />

266). In the study by Ankrapp et al. (277) described above, it<br />

was found that the greatest induction of EGF by estradiol<br />

occurred in the stromal compartment of the gland. Confirmation<br />

of a requirement for ER� in the glandular stroma and<br />

the likely involvement of paracrine growth factors in estrogen<br />

regulation of mammary gland growth has been demonstrated<br />

by Cunha et al. using the tissue recombinant methodology<br />

described previously (see Section III.A.2). Tissue<br />

recombinations of mammary fat pad and epithelium from<br />

wild-type and �ERKO females were constructed and grafted<br />

under the kidney capsule of athymic nude mice for 4 weeks<br />

(278). Extensive ductal growth was observed only in recombinants<br />

involving wild-type (i.e., ER�) stroma, including<br />

those composed of �ERKO ductal epithelium (278). However,<br />

�ERKO (ER��) stroma was unable to induce growth<br />

in the overlying ductal epithelium from either genotype<br />

(278). The authors therefore concluded that stromal ER� is<br />

essential to the mitogenic actions of estradiol in the mam-


382 COUSE AND KORACH Vol. 20, No. 3<br />

mary epithelium, a conclusion also reached from similar<br />

studies in the uterine and vaginal tissue. In light of these data<br />

produced by Ankrapp et al. (277) and Cunha et al. (278),<br />

current studies to determine whether the underdeveloped<br />

�ERKO mammary gland may be rescued by exogenous treatment<br />

with growth factors will prove interesting.<br />

Complete maturation of the mammary gland during pregnancy<br />

involves further ductal growth, extensive branching,<br />

and differentiation of the lobuloalveolar structures along the<br />

ducts, the hallmark of the lactating gland (266). These alveoli<br />

provide for milk production and will eventually fill the remaining<br />

spaces of the fat pad. This stage of mammary gland<br />

development is thought to require the combined actions of<br />

estrogen, progesterone, and PRL (266). The need for progesterone<br />

action in lobuloalveolar development was confirmed<br />

by studies of the PRKO mouse, which exhibited a normal<br />

pubertal ductal structure but was completely refractory to<br />

the induction of lobuloalveolar development after progesterone<br />

treatment (44). A similar phenotype is observed in the<br />

female PRL-receptor knockout (PRLR�/�) mice, which possess<br />

a normal virgin mammary gland as an adult, but a severe<br />

deficit in lobuloalveolar development and lactation after<br />

pregnancy (279). Interestingly, because the homozygous<br />

PRL-receptor (PRLR�/�) knockout females like others are<br />

infertile, this mammary gland phenotype was first detected<br />

in the heterozygous (PRLR�/�) females (279). Furthermore,<br />

the defects in lactation in the PRLR�/� mice were lessened<br />

after multiple pregnancies, due to the compensatory actions<br />

of other pregnancy-associated mammotrophic hormones<br />

that accumulated with each pregnancy (279).<br />

Due to the documented ability of the estradiol-ER� complex<br />

to increase PR expression in various tissues, including<br />

the mammary gland (280), a loss of ER� action may also<br />

result in a loss of PR-mediated progesterone functions. Although<br />

progesterone actions in the mammary gland were<br />

once thought to be strictly differentiative and may actually<br />

oppose the proliferative effects of estrogens, recent work has<br />

suggested that progesterone may also act as a mitogen in the<br />

breast (reviewed in Ref. 181). This is best illustrated by the<br />

less extensive ductal arborization of the mammary gland, as<br />

well as the lack of alveolar structures, in the PRKO mice (44).<br />

Furthermore, in vitro studies have indicated that PR-mediated<br />

progesterone actions can regulate the transcription of<br />

genes for cell cycle-related proteins, growth factors, and<br />

growth factor receptors (181). Therefore, the lack of ductal<br />

growth and differentiation in the �ERKO mammary gland<br />

may be largely due to a lack of physiologically sufficient PR<br />

and subsequent progesterone action. As shown in Table 2,<br />

the average serum progesterone levels in adult female<br />

�ERKOs are 4.0 (�1.1) ng/ml vs. 2.3 (�0.6) ng/ml in wildtype<br />

and are therefore within the normal range of cycling<br />

wild-type females (123). Preliminary analysis by ribonuclease<br />

protection assay indicates that the mammary glands of<br />

adult virgin �ERKO females possess slightly detectable levels<br />

of PR mRNA, but exhibit no increased regulation when<br />

treated with estrogen. Furthermore, current studies in our<br />

laboratory indicate that prolonged progesterone treatment of<br />

adult �ERKO females results in the formation of terminal end<br />

buds and differentiation in the mammary gland, indicating<br />

that the lack of progesterone action caused by disruption of<br />

the ER� gene can be partially overcome with treatment of<br />

superphysiological levels of exogenous progesterone.<br />

A similar secondary effect of targeted disruption of the<br />

ER� gene on the mammary gland may involve a lack of PRL<br />

stimulation. PRL is primarily secreted from the anterior pituitary,<br />

although local synthesis in the mammary epithelium<br />

has also been described (279). This peptide hormone plays a<br />

critical role in mammary gland physiology, especially in the<br />

induction of lobuloalveolar structures and milk production,<br />

as evidenced in the phenotype of the PRLR�/� mice (reviewed<br />

in Ref. 279). However, the synthesis and secretion of<br />

PRL from the anterior pituitary are strongly regulated by<br />

estradiol and ER� (281). Confirmation of the regulatory action<br />

of ER� on the PRL gene is provided by the �ERKO<br />

females, which exhibit a 20-fold decrease in PRL mRNA in<br />

the anterior pituitary (282) and a 5-fold decrease in serum<br />

PRL levels (Table 2) (see Section VI.A.4). Therefore, given the<br />

significant role of PRL in the differentiation to a lactating<br />

mammary gland, it is likely that the phenotype described in<br />

the �ERKO female is also at least partly due to a lack of PRL<br />

stimulation. Support for this hypothesis is provided by current<br />

studies in which elevated serum PRL levels are achieved<br />

in the �ERKO female via wild-type pituitary transplants<br />

placed under the kidney capsule and have resulted in significant<br />

ductal growth and differentiation in the �ERKO host<br />

mammary gland (W. P. Bocchinfuso, J. Lindzey, S. W. Curtis,<br />

J. E. Clark, P. H. Myers, and K. S. Korach, in preparation).<br />

However, preliminary analysis indicates that this partial rescue<br />

of the �ERKO mammary gland by the PRL secreting<br />

wild-type pituitary graft does not occur in an ovariectomized<br />

�ERKO host female, indicating a requirement for ovarian<br />

derived factors as well. Nonetheless, the �ERKO mammary<br />

gland appears to possess the intrinsic tissue components<br />

necessary for pubertal development and pregnancy-induced<br />

maturation, but fails to develop because of the loss of multiple<br />

stimuli that are downstream of ER� action.<br />

B. �ERKO phenotype<br />

Unlike the dramatic underdevelopment observed in the<br />

mammary gland of the �ERKO, no such phenotype is observed<br />

in adult �ERKO females. Virgin �ERKO females of<br />

4–5 months of age exhibit mammary glands that possess a<br />

normal ductal structure that fills the entire fat pad and are<br />

indistinguishable from those of age-matched wild-type females.<br />

This phenotype agrees with our description of minor<br />

amounts of ER� mRNA in the adult mouse mammary gland,<br />

whereas ER� transcripts are easily detectable (Fig. 2) (93).<br />

Furthermore, mammary glands from pregnant and nursing<br />

�ERKO females appear to have undergone normal differentiation<br />

and exhibit the lobuloalveolar structures required<br />

for lactation. Although litter sizes of the �ERKO females<br />

were reduced during the mating study described above, no<br />

marked abnormalities in nursing were observed. Therefore,<br />

the combined data from the �ERKO and �ERKO models<br />

indicate that ER� is the predominant receptor required to<br />

mediate the actions of estrogen in the mammary gland of the<br />

mouse.


June, 1999 ESTROGEN RECEPTOR NULL MICE 383<br />

C. ER� and oncogene-induced tumorigenesis:<br />

Wnt-1/�ERKO mice<br />

Several lines of evidence indicate that breast cancer in<br />

humans is strongly correlated with the extent of lifetime<br />

exposure to estrogen. Most notably, breast cancer almost<br />

exclusively occurs in females and is never seen before puberty<br />

but rather only after several years into the reproductive<br />

life span (283, 284). Furthermore, an increased length of a<br />

woman’s reproductive years, i.e., early menarche and late<br />

menopause, has been associated with an elevated risk of<br />

breast cancer (284), whereas women who have experienced<br />

premature menopause due to natural causes or castration<br />

appear to be at a much lower risk (283). Furthermore, studies<br />

have indicated that women who circulate higher levels of<br />

active estrogens may also be at greater risk of developing<br />

breast cancer (283). In apparent contrast, early pregnancy<br />

tends to provide a protective effect, although it is obviously<br />

associated with an increased level of steroid hormone exposure.<br />

In addition, several years of research have indicated<br />

little positive correlation with the risk of breast cancer and<br />

prolonged use of contraceptive pills composed of synthetic<br />

estrogens and progestins, although this issue remains controversial<br />

(reviewed in Refs. 285 and 286). Still, a large portion<br />

of chemotherapeutics for breast cancer are aimed at<br />

either blocking estrogen action or reducing estrogen levels<br />

(283). Therefore, although an association between estrogen<br />

action and breast cancer is apparent, it involves less understood<br />

yet critical mechanisms, including the periodicity and<br />

cyclicity of hormone exposure as well as the sensitivity of the<br />

end organ (283). This is further complicated by the influence<br />

that environmental exposures, geography, diet, body weight,<br />

and genetics also play in individual risk of developing breast<br />

cancer (283).<br />

Numerous studies have been carried out concerning the<br />

levels of ER and PR in neoplastic breast tissue and the prognostic<br />

value that these parameters may provide (reviewed in<br />

Refs. 287–291). Reports indicate that more than 70% of primary<br />

breast tumors are ER�-positive and exhibit estrogendependent<br />

growth (291). However, the most malignant<br />

mammary tumors are often ER�-negative and exhibit estrogen-independent<br />

aggressive growth, but are thought to<br />

progress from a once ER�-positive cell population (287). In<br />

addition, the role of local aromatase activity and estrogen<br />

production in breast cancer is receiving increased attention<br />

(reviewed in Ref. 292). Added complexity is introduced by<br />

the detection and description of numerous variants of the<br />

ER� transcript in breast cancer tissues, although their possible<br />

role in the etiology of the disease remains speculative<br />

(reviewed in Refs. 28 and 293). Recent reports also described<br />

the detection of ER� transcripts in multiple immortalized<br />

human breast cancer cell lines, normal human breast tissue,<br />

and human breast tumors (64, 73, 101, 102, 294). Vladusic et<br />

al. (73) also characterized an ER� mRNA variant detected in<br />

normal as well as malignant human breast tissue (73).<br />

It is clear from the severely underdeveloped mammary<br />

gland of the �ERKO that estradiol acting via the ER� is a<br />

potent mitogen in the breast. To gain insight into the potential<br />

role of ER� in the induction and promotion of mammary<br />

gland carcinogenesis, we crossed the �ERKO mice with the<br />

MMTV-Wnt-1 mice, a transgenic line that is highly susceptible<br />

to mammary adenocarcinoma. The family of Wnt genes<br />

encode a series of secretory glycoproteins that act in autoand<br />

paracrine pathways to stimulate cell proliferation and<br />

differentiation (reviewed in Ref. 295). The MMTV-Wnt-1<br />

mice possess a transgene designed for targeted overexpression<br />

of the Wnt-1 protooncogene in the mammary gland and<br />

exhibit a nearly 100% and 15% incidence of mammary hyperplasia<br />

and lobuloalveolar adenocarcinoma by 1 yr of age<br />

in females and males, respectively (295, 296). Therefore,<br />

breeding of the two transgenic lines allowed for the generation<br />

of animals that possessed the Wnt-1 transgene on either<br />

a wild-type or �ERKO background and thereby allowed for<br />

the assessment of the role of ER� in the initiation and promotion<br />

of protooncogene-induced mammary tumors (297).<br />

At 6 months of age, virgin wild-type Wnt-1 females exhibit<br />

extensive hyperplasia of the ductal epithelium and aberrant<br />

lobuloalveolar development that occupies the entire fat pad.<br />

This was the expected phenotype based on that described<br />

previously in the original line of Wnt-1 females derived from<br />

a different mouse strain (296). Interestingly, a similar phenotype<br />

of lobuloalveolar hyperplasia was observed in the<br />

rudimentary duct of the �ERKO-Wnt-1 females, although the<br />

extent of ductal growth was much reduced compared with<br />

that seen in the wild-type (269). Nonetheless, the rudimentary<br />

ductal structure previously described in the �ERKO<br />

female was obviously induced to proliferate by the presence<br />

of ectopic Wnt-1 expression. However, comparison of mammary<br />

glands from a series of age-matched animals indicated<br />

that the ductal growth observed in the mammary gland of a<br />

6-month-old �ERKO-Wnt-1 female remained confined to the<br />

nipple region and approximated that seen in a 2.5-month-old<br />

wild-type-Wnt-1 female, illustrating a significant delay in the<br />

proliferation of the �ERKO epithelium (269). Interestingly,<br />

the mammary hyperplasia in the �ERKO-Wnt-1 females did<br />

not markedly progress into the inguinal fat pad, but rather<br />

remained confined to the area of the nipple. Therefore, although<br />

the lobuloalveolar phenotype characteristic of Wnt-1<br />

overexpression was evident, ductal elongation did not occur<br />

in the �ERKO-Wnt-1 female, indicating that the hyperplastic<br />

action of ectopic Wnt-1 expression cannot substitute for ER�mediated<br />

terminal end bud formation and ductal morphogenesis<br />

(297). In the males, Wnt-1-induced epithelial hyperplasia<br />

was obvious in both the wild-type and �ERKO<br />

animals as well, but no distinct difference in growth rates<br />

between the two genotypes was evident (269).<br />

The incidence of lobuloalveolar carcinoma in the Wnt-1<br />

mice mirrored the observations described above for the epithelial<br />

hyperplasia. Wild-type-Wnt-1 and heterozygous-<br />

ER�/Wnt-1 females developed mammary tumors at a rapid<br />

rate, reaching an incidence of 50% by 6 months of age (297).<br />

Ectopic expression of the Wnt-1 gene was also able to induce<br />

tumorigenesis in the �ERKO females and therefore did not<br />

require the presence of functional ER� (297). However, a 50%<br />

incidence in tumors in the �ERKO-Wnt-1 females was not<br />

observed until 12 months of age, twice the time required for<br />

the wild-type-Wnt-1 colony (269). Ribonuclease protection<br />

assays indicated that the level of Wnt-1 transgene expression<br />

was relatively equal in the two ER� genotypes. Prepubertal<br />

ovariectomy had no effect on the overall incidence of tumors


384 COUSE AND KORACH Vol. 20, No. 3<br />

in either genotype, although it resulted in a delayed incidence<br />

of palpable tumors in both, i.e., wild-type-Wnt-1 at<br />

50% at 10.5 momths and a further delay in the �ERKO-Wnt-1<br />

at 50% at 15 months (269). Therefore, ovarian factors clearly<br />

accelerated the growth rate of the Wnt-1-induced adenocarcinoma<br />

in both the wild-type and �ERKO females. However,<br />

multiple pregnancies had no significant effect on the time of<br />

tumor onset in the wild-type-Wnt-1 females (269).<br />

These studies indicate that ER� signaling is not involved<br />

in the induction of hyperplasia and tumorigenesis in the<br />

mammary gland that results from overexpression of the<br />

Wnt-1 protooncogene, but indeed plays a promotional role<br />

in these phenotypes. Similar findings of a promotional role<br />

for the ER� in uterine carcinogenesis were reported in transgenic<br />

mice that overexpressed an ER� gene (298). A possible<br />

explanation for the observed tumor latency may be drawn<br />

from an inherent difference between the wild-type and<br />

�ERKO mammary glands, i.e., because the �ERKO-Wnt-1<br />

glands possess a reduced amount of ductal morphogenesis<br />

and proliferating epithelia due to the lack of ER�, the number<br />

of cells most susceptible to neoplastic transformation may be<br />

decreased (297). Interestingly, the growth rates of the ductal<br />

hyperplasia and ultimate adenocarcinomas were also significantly<br />

reduced when the animals were ovariectomized, even<br />

in those lacking functional ER�. A possible explanation for<br />

this finding may be that ovarian estradiol is acting via non-<br />

ER�-mediated pathways, suggesting a possible role for ER�.<br />

However, in contrast to reports in other species, ER� mRNA<br />

remains difficult to detect by standard assays in the mouse<br />

mammary gland, including the glands of the �ERKO (93) and<br />

Wnt-1 females (297). Furthermore, ovariectomy also results<br />

in the loss of other gonadal hormones in addition to estrogen.<br />

Remarkably, overexpression of the Wnt-1 gene in the<br />

�ERKO mammary gland resulted in significant increases in<br />

PR mRNA levels compared with the low basal levels detected<br />

in control �ERKO glands, when assayed by ribonuclease<br />

protection assay (297). Recently, Shyamala et al. (299)<br />

showed that misregulation of the PR gene resulting in increased<br />

levels of PR A through transgenic methodologies may<br />

be tumorigenic in the mammary gland. Therefore, the ability<br />

of Wnt-1 to compensate for the reduced PR levels generated<br />

by the loss of ER� action may provide a pathway by which<br />

the role of ER� in tumor induction and promotion may be<br />

overridden. In addition, if the PR pathway was involved in<br />

the promotion of the Wnt-1 lobuloalveolar carcinomas,<br />

ovariectomy would be expected to result in a delay in their<br />

growth.<br />

In summary, the �ERKO/Wnt-1 mouse model has demonstrated<br />

that artificially induced hyperplasia and tumorigenesis<br />

of the mammary gland can take place in the absence<br />

of ER� signaling. These findings have potential importance<br />

in understanding the etiology of human breast cancer by<br />

illustrating that oncogenic induction of mammary neoplasia<br />

can occur in an ER�-negative tissue. Extending these observations<br />

further may indicate that hormone-independent<br />

breast tumors can originate from a cell population lacking<br />

ER� as well as those that are ER� positive. The characterization<br />

of mice produced from crosses of other overexpressing<br />

transgenic models, e.g., neu, with the �ERKO are cur-<br />

rently being carried out in a fashion similar to those studies<br />

described above.<br />

V. Reproductive Tract Phenotypes of the Male<br />

More than 40 yr ago, Jost employed a series of classical<br />

organ ablation studies to demonstrate that development of<br />

the male reproductive system was dependent on the secretion<br />

of testosterone and the peptide, anti-Müllerian hormone,<br />

by the fetal testes (107). We now know that development of<br />

the undifferentiated fetal gonad to a testis is determined by<br />

the presence of testis-determining factors or genes localized<br />

to the Y chromosome, e.g., the SRY gene (300). Testosterone<br />

produced from the Leydig cells of the fetal testes is crucial to<br />

the survival of the Wolffian duct, the primordial structure of<br />

the male reproductive tract, and its differentiation into the<br />

male reproductive structures (143, 192). In turn, secretion of<br />

anti-Müllerian hormone from the Sertoli cells of the fetal<br />

testis induces regression of the primordial female reproductive<br />

structures in the developing fetus (143, 192). Differentiation<br />

of the epididymis, ductus deferens, and seminal vesicle<br />

from the Wolffian duct is stimulated by testosterone,<br />

whereas development of the lower structures, the prostate,<br />

bulbourethral glands, urethra, scrotum, and penis, is primarily<br />

influenced by the more potent testosterone metabolite,<br />

5�-dihydrotestosterone (DHT) (143). The importance of<br />

androgen action in the development of the male reproductive<br />

tract is evident by the feminized phenotype that results in<br />

human males with complete androgen insensitivity syndrome<br />

(cAIS), caused by a naturally occurring mutation that<br />

results in the lack of functional AR. The phenotypes of the<br />

cAIS human male and the analogous testicular feminized<br />

male (Tfm) rodent are characterized by the complete absence<br />

of either male or female internal reproductive structures<br />

except for the presence of inguinal testes (37, 301). External<br />

genitalia in the cAIS human and Tfm mouse are indistinguishable<br />

from the normal wild-type female, but a short and<br />

often blunt-ended vagina is present (37, 301). During puberty<br />

and the later stages of sexual maturation in the male, virulization<br />

of the external genitalia is dependent on the actions<br />

of DHT (143, 302). This is illustrated in human males lacking<br />

sufficient 5�-reductase activity and DHT synthesis, who exhibit<br />

normal internal reproductive structures and testosterone<br />

levels but severely undervirulized external structures<br />

(reviewed in Ref. 37). Once again, the critical role of androgens<br />

in the development of the male reproductive tract was<br />

reiterated by the complete lack of gonadal development in<br />

mice homozygous for a disruption of the gene encoding SF-1<br />

(137, 138). Therefore, androgen action is critical to the development<br />

and function of the tissues of the male reproductive<br />

tract from fetal stages through adulthood, whereas a<br />

defined role for estrogen remains unclear.<br />

Although there is no apparent role for estrogen action in<br />

the development of the male reproductive tract, studies have<br />

reported significant effects of early exposure to estrogen<br />

agonists and/or antagonists on the adult male reproductive<br />

system. However, it was generally concluded that such effects<br />

were caused by estrogen actions at the hypothalamicpituitary<br />

level that ultimately resulted in decreased stimu-


June, 1999 ESTROGEN RECEPTOR NULL MICE 385<br />

lation of the testis and reduced androgen production (303,<br />

304). Still, a series of studies by McLachlan et al. demonstrated<br />

that perinatal exposure to the synthetic estrogen,<br />

DES, results in an assortment of apparently direct defects in<br />

the murine male reproductive tract, including undescended<br />

testes, epididymal cysts, aberrant expression of estrogeninducible<br />

genes, adenocarcinoma, and sterility (305–308).<br />

However, these studies indicated a toxic effect of developmental<br />

exposure to pharmacological levels of estrogens<br />

rather than a required physiological function of estrogen in<br />

the male reproductive tract. Investigators have reported the<br />

detection of ER by steroid autoradiography (309, 310) and<br />

immunohistochemistry (112) in the testes and accessory tissues<br />

of the male tract, even as early as day 16 of gestation in<br />

the mouse. In addition, the newly described ER� also appears<br />

in varied amounts in the testis, epididymis, and prostate<br />

of the rodent (93, 121, 311–314) and rhesus monkey (96).<br />

Hence, these studies suggest the presence of a functional<br />

estrogen- signaling system in the male reproductive tract and<br />

further demonstrate the direct detrimental effects that may<br />

occur when this system is aberrantly activated.<br />

A. Testicular function and spermatogenesis<br />

A defined role of ER and estrogen action in the function<br />

and maintenance of the male reproductive tissues remained<br />

elusive until the generation of the �ERKO mice. As expected<br />

in the presence of a functional androgen-signaling system,<br />

the reproductive tract of the �ERKO male mouse undergoes<br />

apparently normal prenatal development to produce internal<br />

and external structures that are indistinguishable from<br />

wild-type littermates. However, at approximately 20 weeks<br />

of age, significant decreases in the weight of the testis and<br />

epididymis/vas deferens are observed, whereas the seminal<br />

vesicle/coagulating glands and prostate appear normal in<br />

size (315, 316). As shown in Table 2, circulating gonadotropins<br />

in mature �ERKO males include levels of FSH within the<br />

wild-type range, whereas LH is slightly elevated (315, 317).<br />

In accordance with the increased LH secretion are observations<br />

of Leydig cell hyperplasia in the testis (121) and serum<br />

testosterone levels of approximately 2-fold in the �ERKO<br />

male compared with age-matched wild-type males (Table 2)<br />

(315, 317). Therefore, the decreased weights reported for the<br />

accessory organs of the male reproductive tract are not a<br />

secondary effect of any phenotype in the hypothalamic-pituitary<br />

axis, but rather a direct result of the loss of ER�mediated<br />

estrogen action within these tissues.<br />

One of the most striking initial observations in the �ERKO<br />

was complete infertility in the males when tested in a continuous<br />

mating study with wild-type, known-fertile females<br />

(315). No difference in fertility was observed in the heterozygous<br />

�ERKO males (315). Although the etiology of such<br />

infertility was unclear, initial experiments in which �ERKO<br />

males were placed in a mating environment with hormoneprimed<br />

wild-type females resulted in significantly fewer copulatory<br />

plugs, indicating a severe deficit in normal sexual<br />

behavior (315, 318). Further investigations have demonstrated<br />

that infertility in the �ERKO male is due to pleiotropic<br />

effects resulting from disruption of the ER� gene. In addition<br />

to a lack of normal sexual behavior, �ERKO male mice ex-<br />

hibit significantly lower sperm counts that further diminish<br />

with age compared with their wild-type and heterozygous<br />

litter mates (315). This is compounded by defects in the<br />

function of those sperm that are produced in the testis of the<br />

�ERKO male, including obvious deficits in motility and a<br />

complete inability to fertilize wild-type oocytes in an in vitro<br />

assay (315).<br />

The testes of the �ERKO male are slightly smaller than<br />

wild-type but do develop normally and possess the usual<br />

complement of seminferous tubules surrounded by interstitial<br />

tissue and Leydig cells (Figs. 6 and 7). Stimulation of the<br />

Leydig cells by LH and subsequent androgen synthesis in the<br />

�ERKO testes appears sufficient as mentioned earlier. However,<br />

a report by Hutson et al. (316) indicated a greater in-<br />

FIG. 6. Testes and seminal vesicle/coagulating gland weights in the<br />

wild-type and �ERKO males. Mean total testes weights (A) and seminal<br />

vesicle/coagulating gland weights (presented as % body weight)<br />

of age-matched wild-type and �ERKO males from 4–22 months of age.<br />

Sample number in each group was �9. Error bars indicate SEM. Statistical<br />

significance is indicated as follows (t test, wild-type vs.<br />

�ERKO, unequal variance): *, P � 0.05; **, P � 0.01; ***, P � 0.001.


386 COUSE AND KORACH Vol. 20, No. 3<br />

FIG. 7. Histology of representative adult �ERKO testis. a, Longitudinal section of a testis and portions of the epididymis from an adult �ERKO<br />

mouse (79 days old). The cranial portion of the testis and part of the caput epididymis (Cp Ep) are in the upper portion of the panel, and the<br />

caudal pole of the testis and cauda epididymis (Ca Ep) are in the lower portion. As indicated, the rete testis (RT) is conspicuously dilated. The<br />

seminiferous tubules in the caudal pole of the testis have a thin seminiferous epithelium and a considerably dilated lumen (as indicated by the<br />

solid arrows), spermatogenesis is disrupted, and fluid is present in the interstitium surrounding these tubules. However, in the seminiferous<br />

tubules in the cranial pole of the testis, the seminiferous epithelium is thicker, the lumen is smaller, spermatogenesis is ongoing, and the tubules<br />

are closely situated. [Reproduced with permission from E. M. Eddy et al.: Endocrinology 137:4796–4805, 1996 (315). © The <strong>Endocrine</strong> Society.]<br />

b, High-power magnification (132�) of seminiferous tubules from an adult �ERKO male (110 days old) illustrating an apparently normal<br />

seminiferous tubule possessing an epithelium (SE) composed of Sertoli cells and spermatogonia (open arrows), juxtaposed with a tubule<br />

possessing a severely dilated lumen and a thinning seminiferous epithelium (solid arrow) with little active spermatogenesis occurring. Leydig<br />

cells (LC) are present in the interstitial space.<br />

cidence of retraction of the testes into the abdomen and a<br />

smaller yet more muscular cremaster sac in �ERKO males<br />

compared with wild-type counterparts. This anomaly was<br />

noticed only after excision and fixation of the urogenital<br />

system and was not observable upon external examination of<br />

the animals. Interestingly, this is one of the few phenotypes<br />

of the �ERKO that has also been observed in the heterozygous<br />

littermates as well (316). This same laboratory had<br />

earlier employed the Tfm mouse, a naturally occurring androgen-resistant<br />

mutant, to produce strong evidence in support<br />

of a biphasic model of testicular descent in which androgens<br />

are critical to the second step of testis migration, i.e.,<br />

from the internal inguinal ring to the scrotum via action on<br />

the genito-inguinal ligament (or gubernaculum) (319). Previous<br />

studies had shown that exposure of fetal male mice to<br />

exogenous estrogens resulted in a failure in testicular descent,<br />

although it was concluded that this was probably an<br />

indirect effect due to inhibition of the hypothalamic-pituitary<br />

axis and anti-Müllerian hormone (319). However, the significant<br />

incidence of undescended or retracted testes in the<br />

�ERKO strongly suggests a previously unrecognized and<br />

possibly direct role for the ER� in development of the male<br />

reproductive tract.<br />

Spermatogenesis, the production of free spermatozoa, is<br />

the primary function of the testes. Similar to the process of<br />

gametogenesis in the females described earlier, spermatogenesis<br />

is an equally complex process involving several different<br />

cell types and hormone actions. This is reiterated by<br />

the severely impaired spermatogenesis observed in naturally<br />

occurring inactivating mutations of the AR gene (38, 301), as<br />

well as in a number of knockout mouse models (reviewed in<br />

Ref. 110), e.g., those for the genes encoding Hsp70–2 (320),<br />

CREM (321, 322), DAZLA (323), and HR6B (324). Although<br />

FSH was previously thought to be essential to spermatogenesis,<br />

mice possessing targeted disruptions for the FSH�-subunit<br />

gene (241) or the FSH receptor gene (242) are fertile and<br />

show only minor detriments in testicular and sperm function.<br />

Estradiol has been thought to play only a minor role if any<br />

in sperm production, although Sertoli cells lining the seminferous<br />

tubules produce estradiol (325) and express detectable<br />

levels of ER (326). Therefore, it was surprising to observe<br />

severe impairments in spermatogenesis at multiple levels in<br />

the �ERKO male. At 8 wk, the earliest age studied, �ERKO<br />

males possess epididymal sperm counts that are approximately<br />

55% of that found in wild-type littermates (315). This<br />

value continues to decrease with age, with �ERKO males<br />

possessing approximately 13% of wild-type sperm counts at


June, 1999 ESTROGEN RECEPTOR NULL MICE 387<br />

16 wk of age (315). Furthermore, the epididymal sperm collected<br />

from the �ERKO males are characterized by significantly<br />

decreased levels of motility and increased incidence of<br />

sperm heads separated from the flagellum (315). Even those<br />

sperm that possessed normal structure and motility were<br />

unable to fertilize wild-type oocytes in an in vitro fertilization<br />

assay (315). Therefore, despite levels of circulating gonadotropins<br />

and androgen within the normal range, disruption of<br />

the ER� gene has resulted in severe impairments in both<br />

spermatogenesis and sperm function.<br />

As shown in Fig. 7, histological analysis of testes from<br />

sexually mature �ERKO males indicated significant atrophy<br />

of the seminiferous epithelium and severe dilation of the<br />

tubule lumen. At 10–20 days of age, no morphological difference<br />

in the testis was apparent when comparing the<br />

�ERKO with wild-type. However, a distinct morphological<br />

phenotype becomes obvious by 40 days of age and<br />

progresses to produce a completely atrophied testis by 150<br />

days in the �ERKO male (315). Accordingly, sperm counts<br />

decrease as the testicular phenotype worsens, although immunohistochemical<br />

detection of Hsp70–2, a germ cell-specific<br />

protein, was possible even in the most severely disrupted<br />

tubules (315).<br />

Further characterization of testes from mature �ERKO<br />

males indicated a prominent rete testis that is dilated and<br />

protrudes into the interior of the organ as well as severely<br />

dilated efferent ductules (Fig. 7) (315, 327). The rete testes are<br />

composed of a network of intercommunicating channels located<br />

in the posterior-cranial portion of the testes and serve<br />

as a pathway by which suspended spermatozoa can pass<br />

from the testis to the epididymis. Connecting the rete testis<br />

to the epididymis are the efferent ducts, a series of multiple<br />

channels thought to play a significant role in reabsorption of<br />

much of the testicular fluid, and therefore act to also concentrate<br />

the sperm (303). Steroid autoradiography has indicated<br />

that the efferent ducts of the mouse possess the highest<br />

concentration of ER compared with any other region of the<br />

excurrent duct system (309). This expression pattern is evident<br />

in the mouse as early as neonatal day 3 (310). Immunohistochemical<br />

and RNA analyses have shown that ER� is<br />

the predominant form of ER in the efferent ducts and the<br />

cranial portion of the epididymis (311, 328), although ER�<br />

mRNA is also detectable (121, 311).<br />

Previous studies employing surgical ligation of the efferent<br />

ductules reported a severe dilation of the seminiferous<br />

tubules similar to that observed in the �ERKO male (329,<br />

330). Based on the similarity of phenotypes, it became clear<br />

that the testicular anomaly observed in the mature �ERKO<br />

male may be the result of a severe imbalance in the fluid<br />

equilibrium. However, was it due to hypersecretion of fluid<br />

from the testis or insufficient reabsorption of fluid by the<br />

epithelial cells lining the efferent ducts, or possibly a combination<br />

of both? Using surgical techniques to inhibit fluid<br />

transport at different points in the excurrent duct system,<br />

Hess et al. (327) demonstrated that the reabsorption abilities<br />

of the efferent ductules in the �ERKO male were lacking and,<br />

in fact, the secretory activity is actually reduced in the<br />

�ERKO testis. Further characterization indicated a reduction<br />

or often a complete lack of endocytotic vesicles and organelles<br />

common to fluid uptake in the epithelial cells lining<br />

the �ERKO efferent ducts (327). This study was the first<br />

report of a direct ER�-mediated estrogen function in the male<br />

reproductive tract. Interestingly, however, was the inability<br />

of the pure antiestrogen, ICI-182,780, to produce a similar<br />

phenotype in wild-type ductal fragments in in vitro experiments<br />

(327). Although the antagonist was able to cause some<br />

loss of fluid absorption in the wild-type ductal fragment, the<br />

resulting phenotype was not nearly as extreme as that observed<br />

in the �ERKO tissue fragments (327). The authors<br />

proposed that perhaps ER� was possibly mediating an agonistic<br />

effect of the ICI-182,780 and thereby may explain the<br />

lack of full corroboration with the in vivo �ERKO phenotype<br />

(327). This hypothesis was based on the work of Paech et al.<br />

(91), which demonstrated that estrogen antagonists, including<br />

ICI-164,384, may function as an agonist when interacting<br />

with AP-1 complexes in vitro. We, as well as others, have<br />

since shown that ER� mRNA expression in the �ERKO male<br />

reproductive tract is not altered, although its function remains<br />

unclear (93, 121). However, the preservation of ER�<br />

expression in the �ERKO strongly indicates that the reabsorption<br />

functions of the efferent ducts are indeed dependent<br />

on the presence of functional ER�. This view is strengthened<br />

by the lack of a similar testicular phenotype in �ERKO male<br />

mice observed at ages as old as 14 months (47).<br />

Interestingly, the luminal swelling, loss of germinal epithelium,<br />

and atrophy in the seminiferous tubules of the<br />

�ERKO testes appeared to commence at the caudal portion<br />

of the organ and progress toward the cranial region as the<br />

animal aged (Fig. 7) (315). This is thought to be due to a<br />

gradual increases in testicular pressure leading to restricted<br />

blood flow as the phenotype in the rete testis worsens and<br />

fluid accumulates within the encapsulated organ (315). The<br />

result of such decreased circulation is likely to become initially<br />

manifested in the less vascularized caudal region of the<br />

testis and eventually advance to affect the whole gonad.<br />

Despite the severe testicular phenotype that occurs in the<br />

�ERKO male with age, younger males do produce viable<br />

sperm. However, the motility and fertilization abilities of<br />

epididymal sperm collected from �ERKO males are severely<br />

compromised. ER (326) as well as P450 arom (331) have been<br />

reported in Sertoli cells and germ cells of the testis, respectively.<br />

Therefore, a loss of ER�-mediated estrogen action in<br />

the Sertoli cell may alter sperm function. It is also known that<br />

spermatozoa entering the epididymis are unable to fertilize,<br />

and undergo a critically active maturation process as they<br />

pass through the epididymal cords (303). Estradiol treatment<br />

of adult male mice has been reported to increase the rate at<br />

which spermatozoa pass through the epididymis (332). Furthermore,<br />

expression of ER� in the mouse epididymis appears<br />

to be highest in the caput epididymis, where sperm<br />

first enter after exiting the testis (309). Reports of ER� expression<br />

in the epididymis indicate an opposite distribution,<br />

i.e. highest levels are found in the cauda epididymis, in both<br />

the rat (311) and mouse (121). This pattern of epididymal ER�<br />

expression is preserved in the �ERKO male (121). Nonetheless,<br />

normal fertility in the �ERKO male indicates that any<br />

actions of estrogen required for sperm maturation and fertilization<br />

capacity appear dependent on the presence of ER�.<br />

The varied phenotypes leading to infertility in the �ERKO<br />

male have provided great insight into the role that ER� plays


388 COUSE AND KORACH Vol. 20, No. 3<br />

in the development and function of the male reproductive<br />

tract. The importance of functional ER� is reiterated by the<br />

lack of phenotypes and apparent full fertility in the �ERKO<br />

male mice. It is certainly possible that there are overlapping<br />

functions of the two ERs and that compensatory mechanisms<br />

have reduced the observed phenotypes compared with those<br />

that might result if both ERs were lacking. The speculated<br />

phenotype of a double ER knockout, i.e., ��ERKO, might be<br />

similar to that reported in males homozygous for a disruption<br />

of the P450 arom gene (ArKO), and therefore lacking the<br />

capabilities to synthesize estradiol. However, intriguing inconsistencies<br />

are apparent when comparing the phenotypes<br />

of the �ERKO male with those of the ArKO. For example, the<br />

male ArKO mice exhibit no defects in fertility, at least at the<br />

younger ages examined (257). The testicular weight in adult<br />

ArKO males is within the normal range, and histology indicates<br />

normal testicular development with no indication of<br />

a phenotype similar to that of the �ERKO males (257). The<br />

possible existence of currently unknown ligands able to stimulate<br />

the ER pathways required for function of the male<br />

reproductive tract and therefore altering the phenotype in<br />

the ArKO must be considered. It is also possible that much<br />

of the actions of ER� in the male reproductive tract, as inferred<br />

from the �ERKO, may be ligand independent. Such a<br />

phenomenon may explain the findings of Hess et al. (327), in<br />

which an estrogen antagonist was unable to completely induce<br />

an �ERKO phenotype in the ligated wild-type efferent<br />

duct segments in vitro. Since the testicular phenotype in the<br />

�ERKO becomes more severe with age, it is possible that a<br />

similar, yet even further, delay in the manifestation of this<br />

anomaly may exist when the ligand is removed; however,<br />

there are currently no reports on older ArKO males.<br />

B. Accessory sex organs<br />

Certain accessory organs of the male reproductive tract<br />

that have not been already discussed, namely the prostate,<br />

bulbourethral glands, coagulating gland, and seminal vesicles,<br />

warrant mention. These glands have no known specific<br />

function other than to secrete components necessary to the<br />

volume of seminal plasma. All four tissues are dependent on<br />

androgen stimulation for growth and maintenance (reviewed<br />

in Ref. 333). However, ER has been detected in each<br />

during various stages of development in the rat (310). Furthermore,<br />

the prostate in various species appears to express<br />

significant mRNA levels for ER� as well as ER� (93, 96, 311,<br />

313). A series of studies by the Prins laboratory have described<br />

the toxic effects of neonatal DES exposure on the<br />

morphology and biochemistry of the rat prostate, including<br />

the regulation of AR, ER�, and ER� (314, 334–336). Nonetheless,<br />

no obvious abnormalities in the development of<br />

these glands have been observed in either the �ERKO or<br />

�ERKO mice (47, 315). However, categorical studies of these<br />

tissues have not been carried out to date on older animals of<br />

either model. One observation in �ERKO males is a significant<br />

increase in weight of the seminal vesicle/coagulating<br />

gland that becomes more apparent with age, as shown in Fig.<br />

6. This phenotype is most likely the result of continued stimulation<br />

of this tissue by the elevated levels of serum androgen<br />

that exist in the �ERKO (Table 2).<br />

VI. Neuroendocrine System<br />

The mammalian brain and pituitary are clearly target organs<br />

of steroid hormone action. Several studies have demonstrated<br />

a wide distribution of receptors for all three sex<br />

steroid hormones throughout the different regions of the<br />

brain (reviewed in Refs. 337 and 338) as well as in the heterogeneous<br />

cell types of the pituitary (reviewed in Ref. 339).<br />

The regulatory actions of the gonadal steroid and peptide<br />

hormones on the hypothalamic-pituitary axis have been<br />

characterized in a number of laboratory models and may be<br />

the most well understood area among a growing list of hypothesized<br />

actions of steroids in the central nervous system.<br />

In contrast, changes in sensorimotor, cognitive, and emotional<br />

functions that occur in many women during periods<br />

of peak steroid secretion in the ovarian cycle remain less well<br />

understood (340, 341). More recently, estrogen action has<br />

received great attention as a possible influential factor in<br />

several nonreproductive-related brain functions, including<br />

learning and memory, cognitive function, and pain sensitivity,<br />

as well as in the pathology of neurodegenerative diseases,<br />

such as Parkinson’s and Alzheimer’s (reviewed in Ref.<br />

342).<br />

Several reviews over the past 20 yr have thoroughly documented<br />

the known actions of steroid hormones in the central<br />

nervous system (see Refs. 337, 343–348). Sexual dimorphism<br />

has been described in multiple regions of the central<br />

nervous system (reviewed in Ref. 343). Analogous to the<br />

reproductive tract, the neuroendocrine system undergoes a<br />

process of sexual differentiation and maturation that is<br />

heavily influenced by the steroid receptor-signaling pathways.<br />

Briefly, sexual maturation of the neuroendocrine system<br />

may be defined as the acquisition of pituitary responsiveness<br />

to hypothalamic factors and ovarian steroids and<br />

the onset of steroid-induced sexual behavior. In turn, differentiation<br />

of the neuroendocrine system is demonstrated by<br />

the unique ability of the female hypothalamus to induce an<br />

LH surge in response to a rise in serum estradiol (346, 348).<br />

The “organizational” or differentiating effects of perinatal<br />

steroids are permanent and manifested as measurable structural<br />

changes or as subtle fixations in the system’s sensitivity<br />

to the “activational” effects of steroids during adulthood<br />

(343).<br />

Studies have indicated that the imprinting mechanisms<br />

controlled by the sex steroids in the brain are likely via<br />

multiple pathways, including the regulation of cell death,<br />

neuronal growth, and synaptogenesis (343, 347, 349). These<br />

effects are generally considered to be genomic events mediated<br />

via the nuclear receptor pathways (337). Indeed, the<br />

nuclear receptors for estrogen, androgen, and progesterone<br />

all have been detected in various amounts in the fetal, neonatal,<br />

and adult brain of several species (88, 337, 343, 350–<br />

351). Furthermore, the discovery of the ER� has introduced<br />

a new level of complexity to the neuroendocrine system as<br />

it relates to estrogen action. Recent studies have demonstrated<br />

an expression pattern for the ER� in the rat brain that<br />

is as broad as that for ER� as reviewed in Refs. 351 and 352.<br />

Studies in the primate have reported similar findings (96,<br />

354). Of particular relevance are the descriptions by Shughrue<br />

et al. (88) of colocalization of ER� immunoreactivity and


June, 1999 ESTROGEN RECEPTOR NULL MICE 389<br />

ER� mRNA in certain regions of the rat brain, including the<br />

medial nucleus of the amygdala and the periventricular preoptic<br />

nucleus. Therefore, with preliminary studies indicating<br />

distinct tissue localization of the two ERs in the reproductive<br />

tract, the brain may be the ideal tissue for the study of<br />

possible transactivational actions of ER�/ER� heterodimers.<br />

It is important to reiterate that studies have indicated a normal<br />

expression pattern for the ER� gene in the hypothalamus<br />

of the �ERKO mouse (93, 352).<br />

Aside from the receptor-mediated genomic actions of sex<br />

steroids that have been so well characterized, the possibility<br />

of nongenomic effects of gonadal hormones and their metabolites<br />

has also received increased attention. A number of<br />

rapid responses to gonadal steroids in various tissues have<br />

been reported and are believed to occur too soon after steroid<br />

exposure to be mediated by the classical mechanism of hormone<br />

nuclear receptors; therefore, they have been termed as<br />

being “nongenomic” (reviewed in Refs. 347, 355–358). These<br />

include the rapid activation of membrane calcium channels<br />

by progesterone in the maturing oocyte and spermatozoa, by<br />

estrogens in myometrial cells, and by androgens in rat osteoblast<br />

cells (347, 355, 356). Descriptions of similar nongenomic<br />

effects of steroids in the neuroendocrine system<br />

include rapid increases in cAMP levels in neurons, modulation<br />

of the GABA-GABA A receptor function, release of<br />

GnRH and dopmamine from nerve terminals, modulation of<br />

oxytocin receptors, and the release of PRL from GH3/B6<br />

pituitary cells (343, 356, 357). Supportive experimental findings<br />

indicate the presence of membrane steroid receptors,<br />

including those for estradiol, in various cell types (359–361).<br />

Evidence that a membrane ER is structurally similar to the<br />

nuclear ER� was provided by Pappas et al. (362) in which<br />

multiple ER�-specific antibodies were shown to detect and<br />

localize ER� immunoreactivity in the cell membrane. Furthermore,<br />

Blaustein describes findings of extranuclear ER�<br />

immunoreactivity in the cytoplasm, dendritic processes, and<br />

axon terminals of neurons and suggests an active role for<br />

these receptors in neurotransmitter release (reviewed in Ref.<br />

338). Recently, Razandi et al. (363) reported the detection of<br />

membrane ER� and ER� receptors in Chinese hamster ovary<br />

(CHO) cells transfected with an expression vector of the<br />

respective receptor cDNA, indicating that the membrane and<br />

nuclear forms of each ER originate from the same transcript<br />

and exhibit similar affinities for estradiol. These studies further<br />

demonstrated that the membrane-bound ERs were G<br />

protein linked and able to elicit a variety of signal transduction<br />

events, including the induction of cell proliferation (363).<br />

In contrast, Gu et al. (364) recently employed the �ERKO<br />

mouse to illustrate that the documented rapid action of estradiol<br />

on kainate-induced currents in the hippocampus occurs<br />

in the absence of a functional ER� gene, nor does ICI-<br />

182,780 have an inhibitory effect, suggesting that ER� is not<br />

involved as well. Therefore, the putative membrane receptor<br />

involved in mediating the neuronal effect of estradiol in the<br />

hippocampus described by Gu et al. appears to be distinct<br />

from the intracellular nuclear form of the ER as well as that<br />

described by Razandi et al. (363). Regardless, the ultimate<br />

function of the nongenomic signaling pathways of the gonadal<br />

steroids in the proper organization and function of the<br />

mammalian brain remains unclear. Therefore, the ERKO mu-<br />

tant mice provide an excellent model to not only study the<br />

role of the nuclear receptors, but also further the investigations<br />

of steroid hormone actions that may be nuclear receptor<br />

independent.<br />

A comprehensive review of the neuroendocrine system is<br />

beyond the scope of this discussion and has been reviewed<br />

in detail elsewhere (337, 347). However, as was expected,<br />

distinct phenotypes in the neuroendocrine system have become<br />

evident in mice after disruption of the ER� gene. The<br />

ultimate consequences of the lack of ER� action in the neuroendocrine<br />

system are manifested in the ovary of the<br />

�ERKO female and as severe deficits in sexual and field<br />

behavior in both sexes of the �ERKO mice. Due to the relatively<br />

short time in which the �ERKO model has been<br />

available for study, no detailed characterizations of possible<br />

phenotypes in the hypothalamic-pituitary axis of this model<br />

have been carried out. Therefore, this section of the review<br />

will concentrate on what is currently known about the<br />

�ERKO, but will attempt to shed light on the possible distinct<br />

roles of both ERs based on the limited observations of the<br />

�ERKO.<br />

A. Hypothalamic-pituitary axis<br />

The hypothalamus may be thought of as the interface<br />

between the central nervous system and the endocrine system,<br />

i.e., the pituitary. The anatomical location of the hypothalamus,<br />

forming the base of the brain and residing just<br />

above the pituitary, is conducive to a function of translating<br />

neuronal signals from the brain into humoral factors that<br />

stimulate the appropriate actions in the anterior pituitary<br />

(365). The two components are connected by the hypothalamo-hypophyseal<br />

portal system, within which blood<br />

flows predominantly from the hypothalamus to the anterior<br />

pituitary, carrying the appropriate hormonal factors (365).<br />

These hormones act as releasing or inhibiting factors to control<br />

the secretory activity of the pituitary. In contrast, the<br />

posterior pituitary is connected directly to the hypothalamus<br />

via neurons passing through the pituitary stalk and functions<br />

as a storage organ for the hypothalamic hormones, oxytocin<br />

and vasopressin.<br />

The anterior pituitary is composed of at least five distinct<br />

cell types, all derived from a common primordium, which<br />

have been categorized by the particular peptide hormone<br />

they produce and secrete. These cell types are as follows,<br />

with the secretory hormone in parentheses: gonadotrophs<br />

(FSH and LH), corticotrophs (ACTH), thyrotrophs (TSH),<br />

somatotrophs (GH), and lactotrophs (PRL). Early studies<br />

employing steroid autoradiography demonstrated estrogen<br />

binding to varied degrees throughout the different cells of<br />

the anterior pituitary, although discrepancies among species<br />

are evident (reviewed in Ref. 339). These studies have been<br />

followed by those using in situ hybridization and immunohistochemistry,<br />

indicating that the majority of estradiol binding<br />

in the anterior pituitary is due to the expression of ER�<br />

(366, 367). In most species described, gonadotrophs and lactotrophs<br />

exhibit the greatest level of ER� followed by lower<br />

and varied levels of localization to the other cell types (339).<br />

Furthermore, Shupnik et al. (29, 368) have described in the rat<br />

pituitary a number of ER� mRNA isoforms characterized by


390 COUSE AND KORACH Vol. 20, No. 3<br />

the removal of single or multiple exons as well as 5�-sequences<br />

that exhibit little homology to the full-length wildtype<br />

ER�. Although a distinct function for these ER� isoforms<br />

has not yet been determined, their transcription as well<br />

as that of the full-length wild-type ER� gene appears to be<br />

regulated by ovarian steroids (368, 369). Further complexity<br />

has been introduced by several recent descriptions of the<br />

presence of ER� mRNA in the anterior pituitary of the human<br />

(100), monkey (96), and rat (70, 98, 99). Wilson et al. (98)<br />

report that ER� mRNA levels were easily detectable in the<br />

pituitary of the 15-day-old rat, exceeding the levels of ER�<br />

mRNA. These levels decreased in the adult, whereupon a<br />

distinct sex difference became evident in which the anterior<br />

pituitary of the female expressed much higher levels of ER�<br />

mRNA (98). Still, previous studies employing radiolabeled<br />

ligand do not corroborate this difference in the rat anterior<br />

pituitary between the sexes at the level of estradiol binding<br />

(370, 371). Furthermore, there are contrasting reports concerning<br />

the cellular localization of the ER� transcripts in the<br />

rat anterior pituitary. Mitchner et al. (99) reported varied<br />

levels of ER� mRNA throughout the different cell types of<br />

the anterior pituitary. In contrast, Wilson et al. (98) describe<br />

the detection of both ER� and ER� mRNAs in the gonadotrophs,<br />

whereas the lactotrophs appear to possess ER� only.<br />

Nonetheless, as previously mentioned, we find low to undetectable<br />

levels of ER� mRNA in the pituitary of the adult<br />

mouse, including that of the �ERKO (93). However, in light<br />

of the above studies described in the prepubertal rat, which<br />

indicate a possible developmental role for the ER�, similar<br />

assays in the mouse are warranted.<br />

Although the rate of gonadotropin secretion from the anterior<br />

pituitary is directly modulated by the hypothalamus,<br />

the level of circulating sex steroids is the most important<br />

physiological determinant of serum gonadotropin levels in<br />

animals and humans (reviewed in Ref. 251). The positive and<br />

negative regulatory loops of the hypothalamic-pituitary-gonadal<br />

axis have been reviewed in detail and will be summarized<br />

briefly here (251, 362, 372). In short, the hypothalamus<br />

stimulates the synthesis and subsequent secretion of<br />

the gonadotropins into the circulatory system, which then act<br />

to induce gametogenesis and steroidogenesis in the gonads.<br />

The functional gonadotropins, FSH and LH, are composed of<br />

dimers of the common �-glycoprotein subunit (�GSU), with<br />

a distinct �-subunit that confers specificity to the hormone,<br />

i.e., active FSH consists of a FSH-� (FSH�)/�GSU dimer,<br />

whereas LH consists of a LH-� (LH�)/�GSU dimer (373).<br />

Hypothalamic stimulation of the anterior pituitary is via the<br />

release of GnRH, a deca-peptide that functions to positively<br />

regulate the synthesis and secretion of the gonadotropins<br />

from the anterior pituitary (365). A reciprocal hypothalamic<br />

factor that may act to inhibit pituitary secretion of gonadotropins<br />

has not yet been found. Hypothalamic secretion of<br />

GnRH is not tonic but rather in the form of pulses, driven by<br />

a poorly understood oscillating pulse generator (374, 375).<br />

The pulsatile stimulation of the anterior pituitary by GnRH<br />

thereby results in pulsatile gonadotropin release, which has<br />

been shown to be necessary for gonadal function and reproductive<br />

success (374, 375). Gonadotropin stimulation of the<br />

gonads subsequently results in gametogenesis and the synthesis<br />

of gonadal steroid and peptide hormones, which then<br />

feed back to the hypothalamus and pituitary to regulate FSH<br />

and LH secretion (251). However, differences in the system<br />

of feedback loops are apparent between the sexes. While<br />

there exists a tonic system to maintain a constant level of<br />

gonadotropins in both sexes, the female has been endowed<br />

with the ability to produce a surge of gonadotropin secretion<br />

to produce transient levels of hormone that are several fold<br />

higher than the tonic level (365). This massive gonadotropin<br />

surge provides for the reproductive cycle in the female and<br />

is critical to ovulation.<br />

1. Female-negative gonadotropin regulation. There is ample experimental<br />

evidence in several species that estradiol can suppress<br />

the secretion of gonadotropins from the anterior pituitary<br />

(reviewed in Refs. 251, 372, 373, 376, and 377).<br />

Ovariectomy of the female rodent is known to result in significant<br />

elevations in serum FSH and LH that can be returned<br />

to intact levels with physiological treatments of estradiol<br />

(376, 378, 379). The effects of ovariectomy are mirrored in the<br />

gonadotrophs of the anterior pituitary, which exhibit equally<br />

elevated mRNA levels for the gonadotropin subunit genes<br />

(373, 376). Studies in the female rat indicate that by 21 days<br />

post ovariectomy, LH� mRNA levels rise to as high as 20fold,<br />

whereas the increases in FSH� and �GSU mRNA levels<br />

plateau at 4- to 5-fold (373). Daily treatments with estradiol<br />

will return the gonadotropin subunit mRNAs to the pregonadectomy<br />

levels within 7 days in the rat (373, 376).<br />

The ability of estradiol to maintain a tonic level of gonadotropins<br />

via regulation of both the expression of the gonadtropin<br />

subunit genes and the ultimate secretion of the<br />

peptide hormones is well accepted. However, the precise site<br />

at which the steroid exerts this effect has been difficult to<br />

ascertain. This is partly due to the fact that the ER has been<br />

detected in both the hypothalamic regions controlling pituitary<br />

function as well as the anterior pituitary itself, as previously<br />

discussed. Although there is evidence to support a<br />

direct effect of estradiol on both components of the neuroendocrine<br />

axis, it is believed that the hypothalamus may be the<br />

primary site of action in the negative feedback actions (reviewed<br />

in Refs. 251, 372, 373, and 376). In the laboratory<br />

animal, it is believed that castration principally results in an<br />

increased frequency of GnRH pulses and therefore increased<br />

tonic levels of gonadotropins, both of which can be restored<br />

to normal with exogenous estradiol replacement (373, 380–<br />

383).<br />

Numerous studies have produced data to indicate that<br />

estrogen regulation of hypothalamic GnRH secretion may be<br />

the predominant pathway by which transcription of the gonadotropin<br />

subunit genes and gonadotropin secretion is<br />

maintained (373, 376). The most prominent include the characterization<br />

of transgenic mice that possess, within the anterior<br />

pituitary, a measurable reporter gene under the regulation<br />

of a gonadotropin subunit gene promoter, including<br />

the promotor of the human �GSU gene (384), the rat �GSU<br />

gene (385), and the bovine LH� gene (386). As expected,<br />

ovariectomy of the transgenic females resulted in increased<br />

transcription and activity of the transgene reporter gene. Keri<br />

et al. (386) illustrated that estradiol was able to reduce the<br />

postovariectomy rise in transcription of the reporter gene<br />

despite the lack of detectable DNA binding of the ER� to the


June, 1999 ESTROGEN RECEPTOR NULL MICE 391<br />

gonadotropin promoter sequences of the construct. Furthermore,<br />

the postovariectomy rise in promoter activity could be<br />

prevented by administration of a GnRH antagonist to the<br />

transgenic animal, thereby inhibiting GnRH action at the<br />

gonadotrope. Therefore, a loss of estradiol action via ovariectomy<br />

appeared to result in increased GnRH release from<br />

the hypothalamus, which in turn stimulated increased transcription<br />

of the reporter constructs. However, the possibility<br />

of estradiol actions at the level of the gonadotrope that may<br />

directly effect gene transcription or alter GnRH responsiveness<br />

can not be ruled out.<br />

Regardless of the precise site at which estrogens negatively<br />

regulate gonadotropin expression and release, genetic disruption<br />

of the ER� gene was expected to have effects in the<br />

female hypothalamic-pituitary axis that mimic ovariectomy.<br />

Northern blot analysis of RNA from pituitaries of intact<br />

wild-type and �ERKO females indicates this to be true.<br />

Scully et al. (282) demonstrated that in the �ERKO female, the<br />

levels of the �GSU transcript are elevated 4-fold, whereas<br />

FSH� and LH� mRNAs are as high as 7-fold compared with<br />

wild-type. Ovariectomy in wild-type female littermates produced<br />

elevated gonadotropin subunit mRNAs that approximated<br />

the levels observed in the intact �ERKO, indicating<br />

that the effects of an acute loss of estrogen action are similar<br />

to those produced by a hereditary loss of ER� (282). Therefore,<br />

despite the fact that the hypothalamic-pituitary axis of<br />

the �ERKO female is chronically exposed to elevated levels<br />

of estradiol, the significantly increased level of all three gonadotropin<br />

subunit transcripts resemble those of an ovariectomized<br />

female. These data provide strong support for a<br />

critical role of ER�, rather than ER�, in the negative regulation<br />

of transcription of the gonadotropin subunit genes.<br />

However, at this time, studies of the �ERKO have not provided<br />

data to further elucidate the precise mechanism or site<br />

at which a loss in ER� action has resulted in this effect.<br />

Although the transcript levels of both the LH�, FSH�, and<br />

�GSU are significantly elevated in the �ERKO pituitary, this<br />

effect does not extend to the serum levels of the gonadotropins.<br />

Whereas serum LH is elevated 4- to 7-fold in the adult<br />

�ERKO female, levels of FSH appear to be within the normal<br />

range (Table 2) (252). A similar effect is reported in the PRKO<br />

female mice, although the serum LH levels in this mutant<br />

female are not nearly as elevated as those in the �ERKO<br />

female (387). In addition, ovariectomy in the PRKO female<br />

results in a further elevation of serum LH (387), most likely<br />

due to the loss of serum estrogens. This effect is not observed<br />

in the �ERKO female (252), indicating that estradiol is the<br />

predominant steroid hormone maintaining tonic levels of LH<br />

in the female.<br />

As shown in Table 2, the serum gonadotropin levels in the<br />

�ERKO female indicate that only LH is significantly elevated,<br />

whereas FSH is within the wild-type range. This is in<br />

contrast to the levels of FSH� mRNA in the anterior pituitary<br />

of the �ERKO, which are elevated 7-fold and equal to those<br />

exhibited by an ovariectomized wild-type female. Furthermore,<br />

assays of pituitary homogenates from intact �ERKO<br />

females for FSH protein indicate levels within the wild-type<br />

range, suggesting that the divergence between gene expression<br />

and serum levels for FSH does not appear to be due to<br />

a decreased secretory rate of the hormone but rather at the<br />

level of translation. This is in contrast to the ArKO female<br />

mouse, in which serum levels of both gonadotropins are<br />

reportedly elevated 3-fold compared with the wild-type<br />

(257).<br />

There are a number of possible explanations for this observation<br />

in the �ERKO female. Whereas estradiol treatment<br />

of ovariectomized rats has been reported to completely block<br />

the expected increases in LH� mRNA and LH secretion, it<br />

appears to be only partially effective in reducing transcription<br />

of the FSH� gene and secretion of FSH (376). It is now<br />

known that FSH� gene expression and FSH secretion is selectively<br />

regulated in a positive or negative nature by the<br />

peptides activin and inhibin, respectively (reviewed in Ref.<br />

388). This is illustrated in the knockout mouse model of the<br />

activin receptor type II gene, which exhibits suppressed FSH<br />

levels in the adult of both sexes, supporting a role for activin<br />

in the positive regulation of FSH secretion (248). Although it<br />

is believed that the reciprocal effects of the activin/inhibin<br />

peptides are mediated at the level of the anterior pituitary,<br />

the precise mechanisms of action remain unclear. Inhibin has<br />

been shown to alter the levels of GnRH receptor (389) and<br />

decrease FSH� mRNA levels (390–392), as well as inhibit<br />

translation of FSH� mRNA (391, 393). Activin appears to<br />

utilize similar mechanisms to exert opposite effects on FSH�<br />

mRNA transcription and translation and ultimate secretion<br />

of the hormone (394, 395). Therefore, the normal levels of<br />

FSH in the pituitary and serum of female �ERKO mice may<br />

indicate that disruption of the ER� gene has no effect on the<br />

pattern of activin and inhibin secretion. This is supported by<br />

the apparent negative regulation of FSH� at the translational<br />

level in the �ERKO female, suggesting the presence of an<br />

active inhibin-signaling pathway. Further support is provided<br />

by studies indicating that ovariectomy of the �ERKO<br />

female, and therefore a loss of ovarian inhibin secretion,<br />

results in elevated levels of serum FSH similar to those seen<br />

in ovariectomized wild types (252). Estradiol replacement<br />

was partially effective in reducing the serum FSH in the<br />

ovariectomized wild-type but completely ineffective in the<br />

ovariectomized �ERKO (252). These studies provide for at<br />

least two conclusions: 1) the partial effectiveness of estradiol<br />

in reducing serum FSH levels in the wild-type was likely via<br />

functional ER� in the hypothalamus that may complement<br />

the actions of inhibin in the intact female; and 2) ovarian<br />

factors other than estradiol, most likely inhibin, are maintaining<br />

normal serum FSH levels in the �ERKO female, possibly<br />

by mechanisms that override the loss of ER� action. It<br />

is also possible that both activin and inhibin synthesis and<br />

secretion may be altered in the �ERKO female, but do not<br />

result in a net difference in serum FSH levels. Future investigations<br />

to determine the serum levels of the activin/inhibin<br />

subunit peptides and their functional dimers in the �ERKO<br />

female are warranted.<br />

Another possible explanation for the selective increase in<br />

serum LH in the �ERKO female may be that a lack of ER�<br />

action has resulted in aberrations in the GnRH pulse frequency<br />

and amplitude that are more conducive to LH secretion<br />

from the anterior pituitary. Normal levels of �GSU<br />

transcripts can be maintained with constant GnRH stimulation<br />

of the anterior pituitary (376). However, normal expression<br />

of both gonadotropin �-subunit genes and gonadotro-


392 COUSE AND KORACH Vol. 20, No. 3<br />

pin secretion requires pulsatile stimulation from the<br />

hypothalamus, and each responds differently depending on<br />

the amplitude and frequency of GnRH stimulation (373, 376).<br />

Varied expression of the GnRH receptor on the cell surface<br />

of the gonadotrophs also varies with the level of hypothalamic<br />

stimulation (396). This mechanism, by which the level<br />

GnRH receptors can be differentially regulated and thereby<br />

modify the gonadotropin responsiveness to GnRH, has been<br />

proposed to be a key element in the differential regulation of<br />

the two different gonadotropins by the same releasing hormone<br />

(397). Studies in the rat have revealed that more rapid<br />

GnRH pulses (15–60 min) favor the secretion of LH, whereas<br />

slower pulses (120 min) allow for secretion of FSH (251, 376).<br />

Positive regulation of LH synthesis and secretion is also more<br />

sensitive to the amplitude of GnRH stimulation (251, 376).<br />

Therefore, a loss of ER� action during development and<br />

maturation of the hypothalamic-pituitary axis may have resulted<br />

in a pattern of GnRH secretion that favors the translation<br />

and secretion of LH rather than FSH. The influence that<br />

ER� may have, including a possible compensatory role in the<br />

�ERKO hypothalamus, remains to be evaluated.<br />

2. Female-positive gonadotropin regulation. In addition to maintaining<br />

tonic levels of serum gonadotropins, estradiol also<br />

plays a central role in the preovulatory gonadotropin surge<br />

mode in the female (reviewed in Refs. 263, 264, and 377).<br />

Differentiation of the neuroendocrine system results in the<br />

development of mechanisms necessary to produce a dramatic<br />

preovulatory rise in serum gonadotropins in response<br />

to the positive feedback of ovarian steroids. In the rodent,<br />

developmental differentiation of this pathway in the neuroendocrine<br />

system is unique to the female (365). The surge<br />

in serum LH and FSH is the hallmark of the female cycle and<br />

is critical to ovulation as well as the synchronized induction<br />

of appropriate sexual behavior and, therefore, is vital to the<br />

female ovarian cycle and fertility.<br />

Numerous studies have demonstrated that estradiol is<br />

required for the preovulatory gonadotropin surge, and that<br />

the timing and dose of estradiol exposure may be the most<br />

critical parameters (reviewed in Ref. 377). As with the negative<br />

regulatory effects of estradiol, the precise site of action<br />

for the sex steroids during the preovulatory gonadotropin<br />

surge also remains unclear. However, it is believed to be the<br />

result of the combined effects of external and internal cues<br />

transduced from the brain and the positive feedback of gonadal<br />

hormones that provide for a synchronized pattern of<br />

GnRH secretion upon an anterior pituitary that has been<br />

rendered transiently hypersensitive to the releasing hormone<br />

(reviewed in Ref. 377). In the monkey, destruction of the<br />

neurons involved in GnRH production can be overcome with<br />

pulsatile administration of exogenous GnRH, whereupon a<br />

gonadotropin surge can be produced with exogenous estradiol,<br />

indicating the pituitary as the predominant factor in the<br />

surge (398). However, this is not possible in the rodent,<br />

apparently due to a greater level of interdependency among<br />

the components of the neuroendocrine axis (365).<br />

Therefore, although the rises in ovarian estrogen secretion<br />

are critical to the generation of a gonadotropin surge, the<br />

pathways involved are poorly understood. A number of<br />

mechanisms involving direct actions of estrogen in the brain<br />

have been proposed and recently reviewed (264). However,<br />

it is unlikely that the actions of estrogen are via direct interaction<br />

with GnRH neurons since these processes appear to<br />

be devoid of ER (264). Therefore, the actions of estradiol<br />

appear to result in indirect stimulation of the hypothalamic<br />

neurons that synthesize and release GnRH. Possible mechanisms<br />

include steroid interaction with receptor-positive<br />

monoaminergic and opoid neurons that may mediate the<br />

ultimate effects to GnRH neurons, possibly via modifications<br />

in the levels of catacholamines, glutamate, �-aminobutyric<br />

acid, neuropeptide Y, �-endorphins, and galanin (reviewed<br />

in Refs. 263 and 264). At the level of the anterior pituitary, the<br />

preovulatory increases in estradiol may act in concert with<br />

GnRH to enhance gonadotrope sensitivity to the forthcoming<br />

rise in releasing hormone by increasing the levels of GnRH<br />

receptor (263, 377). Furthermore, the 5�-flanking region of the<br />

rat LH� gene has been shown to possess an imperfect estrogen-responsive<br />

element that is able to bind ER� and confer<br />

estrogen responsiveness to a chimeric promoter-reporter<br />

gene construct in vitro (399). Therefore, the estradiol-ER�<br />

complex may also act to directly increase LH� mRNA levels<br />

before the LH surge (251).<br />

Attempts to elicit an LH surge in the ovariectomized<br />

model with acute estradiol treatments have been reported to<br />

be only partially effective, indicating that ovarian factors<br />

other than estradiol are also required for a full physiological<br />

response (reviewed in Ref. 263). It is now known that the<br />

actions of progesterone and the PR are also a necessary<br />

component in the induction of the gonadotropin surge (reviewed<br />

in Ref. 263). The role of progesterone and PR in<br />

facilitating the preovulatory surge may be to induce a rapid<br />

release of GnRH from the hypothalamus, as well as possibly<br />

mediate a decrease in ER levels in the anterior pituitary,<br />

thereby possibly counteracting the inhibitory effects of estradiol<br />

(263). Although the precise mechanism of action may<br />

be unclear, the complete lack of a preovulatory surge in intact<br />

PRKO female mice provides strong support for the requirement<br />

of this steroid receptor (387).<br />

A cooperative role between the estrogen- and progestronesignaling<br />

pathways in the induction of the preovulatory<br />

surge may include the ability of estradiol to stimulate increased<br />

PR levels in both the hypothalamus and anterior<br />

pituitary, thereby increasing the sensitivity of these tissues to<br />

progesterone (400, 401). Shughrue et al. have shown that<br />

estradiol-induced increases in PR expression in the preoptic<br />

nucleus are possible in the �ERKO female (Fig. 8) and suggest<br />

that this may be a compensatory action of ER� (400).<br />

These same studies demonstrated that the preoptic nucleus<br />

of the hypothalamus in intact �ERKO females possessed a<br />

significantly greater level of PR mRNA when compared with<br />

wild-types, perhaps due to chronic stimulation of ER� by the<br />

elevated serum estradiol (400). Evidence to support this hypothesis<br />

is the apparent decrease in the level of PR transcripts<br />

observed after ovariectomy in the �ERKO, which can be<br />

returned to intact levels 6 h after a single treatment with<br />

estradiol (Fig. 8) (400).<br />

The preoptic area of the hypothalamus, more specifically,<br />

the anteroventral periventricular nucleus of the preoptic region<br />

(AVPV), is thought to play a critical role in transducing<br />

the gonadotropin surge via interactions with the GnRH neu-


June, 1999 ESTROGEN RECEPTOR NULL MICE 393<br />

FIG. 8.In situ hybridization for progesterone receptor (PR) mRNA in female WT and �ERKO hypothalamus. A, PR mRNA was detected in the<br />

medial preoptic nucleus of wild-type (a) and �ERKO females (b) 5 days afer ovariectomy. Also shown is the increased detection of PR mRNA<br />

in ovariectomized wild-type (c) and �ERKO (d) females 6 h after treatment with 5 �g of estradiol. Asterisks indicate the third ventricle. B,<br />

Quantitative analysis of the hybridization signal shown in panel A. Note the dramatic increase in PR hybridization signal when ovariectomized<br />

(OVX) wild-type mice were treated with estradiol (E 2). Similarly, the hybridization signal seen in intact �ERKO females is attenuated by<br />

ovariectomy, but augmented to intact levels when ovariectomized females were treated with estradiol. Statistical significance is indicated as<br />

follows: **, P � 0.01, ***, P � 0.001. [Reproduced with permission from P. J. Shughrue et al.: Proc Natl Acad Sci USA 94:11008–11012, 1997<br />

(400). © National Academy of Sciences, USA]<br />

rons. Unlike most sexually dimorphic nuclei, the AVPV is<br />

actually larger and composed of a greater number of dopaminergic<br />

neurons in the female compared with the male<br />

(402). In male rodents, this region is rendered inoperative by<br />

the actions of testosterone during differentiation (365), an<br />

effect that can be reproduced in females with neonatal testosterone<br />

or estradiol exposure (403, 404). Therefore, masculinization<br />

of this portion of the brain involves the destruction<br />

of a large portion of these neurons and is believed to be<br />

due to local aromatization of testosterone to estradiol and<br />

subsequent activation of ER-mediated pathways (405). In<br />

support of this hypothesis, Simerly et al. (405) reported that<br />

the AVPV region of �ERKO males possess a population of<br />

dopaminergic neurons more characteristic of a wild-type<br />

female, confirming a critical role of ER� in this differentiation<br />

process. Furthermore, the numbers of dopaminergic neurons<br />

in the female �ERKO are only slightly reduced when compared<br />

with wild-type, indicating a morphologically normal<br />

AVPV region (405). Therefore, with the �ERKO female exhibiting<br />

an apparent preservation of estrogen-induced increases<br />

in hypothalamic PR and a wild-type-like female phenotype<br />

in the AVPV region, it is conceivable that the<br />

hypothalamic mechanisms required for induction of the preovulatory<br />

surge may be intact.<br />

3. Males: gonadotropin regulation. Because of the more prominent<br />

role of testosterone in the male, certain issues specific<br />

to the male hypothalamic-pituitary axis are worthy of discussion.<br />

Of course, lower aromatase activity in the testis<br />

results in circulating levels of estradiol in the male that do not<br />

approach those observed in the intact cycling female. Therefore,<br />

it would be expected that distinct mechanisms of steroid<br />

feedback and regulation of gonadotropin synthesis and secretion<br />

from the hypothalamic-pituitary axis have evolved in<br />

males, presumably one likely to be more dependent on tes-<br />

tosterone. This difference is thought to occur at the level of<br />

the hypothalamus since the anterior pituitary generally exhibits<br />

no sexual differentiation (365) and possesses receptors<br />

for all sex steroids (339).<br />

A critical role of testosterone and AR-mediated actions in<br />

the negative regulation of gonadotropin secretion in the male<br />

is illustrated by the elevated serum LH levels in Tfm mice<br />

(301) and in humans with androgen insensitivity syndromes<br />

(38). As in the female, transcription of the gonadotropin<br />

subunit genes is significantly elevated after castration in the<br />

male, although peak levels are reached much earlier (�7<br />

days) (373). Furthermore, FSH� mRNA levels appear to return<br />

to precastration levels by 28 days, whereas the levels of<br />

LH� and �GSU transcripts remain elevated (373). Estradiol<br />

is equally effective as testosterone in reducing serum LH<br />

levels that result after castration in the male (reviewed in Ref.<br />

373). These data, along with the documented presence of<br />

P450 arom activity (reviewed in Refs. 406 and 407) and wide<br />

distribution of ER in the hypothalamic-pituitary axis support<br />

a role for locally synthesized estradiol and ER action in male<br />

gonadotropin synthesis and/or secretion (88, 339). Furthermore,<br />

adult male ArKO mice exhibit elevated levels of serum<br />

LH despite possessing significantly high circulating testosterone<br />

(257). Therefore, the roles of estradiol and testosterone<br />

often appear overlapping as well as distinct, making obvious<br />

the complexity of the steroid-feedback mechanisms that exist<br />

in the male.<br />

Any specific role the ER� may play in the regulation of<br />

gonadotropin synthesis and secretion in the male would be<br />

expected to become apparent in the �ERKO. Adult �ERKO<br />

males exhibit levels of hypothalamic GnRH, pituitary FSH�<br />

mRNA, and serum FSH that are within the normal range<br />

when compared with wild-type littermates (Table 2) (317).<br />

The normal levels of FSH� mRNA in the pituitary of �ERKO


394 COUSE AND KORACH Vol. 20, No. 3<br />

males are in stark contrast to the significantly elevated levels<br />

found in the female �ERKO mice (282). This contrast between<br />

the sexes may reflect differences in the inhibin/activin levels<br />

or may represent a definitive sexual differentiation in the<br />

transcriptional regulation of the FSH� gene in the mouse,<br />

indicating that androgens are the primary acting steroids in<br />

the male. However, although not as extreme as those found<br />

in the �ERKO female, pituitary LH� mRNA and serum LH<br />

levels are increased 2-fold in the adult �ERKO male (Table<br />

2) (317).<br />

In a series of experiments, Lindzey et al. (317) demonstrated<br />

that castration results in the expected elevated levels<br />

of serum LH in the wild-type, and a further increase in the<br />

already elevated LH levels in �ERKO males. The rise in<br />

serum LH that occurs upon castration even in the �ERKO<br />

male suggests that either estradiol-ER� or androgen-mediated<br />

mechanisms are maintaining the lower LH levels in the<br />

intact animal. Once again however, the mouse pituitary (including<br />

the �ERKO) appears to possess very little if any ER�<br />

mRNA (93), although ER� is expressed normally in the hypothalamic<br />

regions in the �ERKO (93, 352). Estradiol treatment<br />

of castrated animals over a period of 3 weeks reduced<br />

the serum LH levels to normal in the wild-type males,<br />

whereas no effect was observed in the �ERKO, indicating a<br />

requirement for ER� in this process (317). Although treatments<br />

of similar castrated males with testosterone was completely<br />

effective in producing an inhibitory effect on LH<br />

release in the wild-types, it was only partially effective in the<br />

�ERKO (317). The authors therefore concluded that the ability<br />

of testosterone to fully restore normal levels of LH in the<br />

sera of castrate wild-type males but only partially in the<br />

�ERKO males suggests that local aromatization of testosterone<br />

to estradiol and subsequent activation of ER�-mediated<br />

pathways act to enhance the negative feedback effects of<br />

androgens in the male hypothalamic-pituitary axis (317).<br />

However, the inability of testosterone to completely suppress<br />

the serum LH in the �ERKO male may be related to the<br />

dosage used in these studies. Strong evidence of AR-dependent<br />

regulation of LH secretion in the male �ERKO is found<br />

in preliminary experiments in which treatment with an antiandrogen<br />

(flutamide) increased serum LH by 3- to 10-fold<br />

in wild-type and �ERKO males, respectively. This suggests<br />

that �ERKO males have come to rely entirely on AR-mediated<br />

actions to regulate LH secretion, whereas the ER� continues<br />

to play a role in the wild-type.<br />

In these same studies, Lindzey et al. (317) illustrated that<br />

prolonged treatment with DHT, the more potent and nonaromatizable<br />

androgen, resulted in no reduction in the castrate<br />

levels of serum LH in wild-type but was partially effective<br />

in the �ERKO male. However, the DHT was effective<br />

in restoring hypothalamic GnRH content levels to normal in<br />

castrate males of both genotypes (317). Therefore, the enhanced<br />

effect of DHT in negatively affecting the hypothalamic-pituitary<br />

regulation of serum LH, including the inhibition<br />

of hypothalamic GnRH release, remains a puzzling<br />

phenomenom unique to the �ERKO male. It is possible that<br />

a lack of ER� action during development resulted in a “reorganization”<br />

of the hypothalamic-pituitary axis in the<br />

�ERKO male, and thereby somehow allowed for an increased<br />

sensitivity to androgens (317). Further studies in the<br />

�ERKO as well as the �ERKO males may help elucidate these<br />

unexpected results.<br />

4. PRL regulation. PRL possesses more biological actions than<br />

all of the other anterior pituitary hormones combined. A<br />

recent review by Bole-Feysot et al. (279) thoroughly covered<br />

the current knowledge of the diverse actions of PRL, including<br />

its functions as a hormone, growth factor, neurotransmitter,<br />

and immunoregulator. A reflection of the multiple<br />

functions of PRL is the equally broad distribution of PRLbinding<br />

sites throughout the many physiological systems in<br />

vertebrates (279). The well known effects of PRL in reproduction<br />

include a critical role in the differentiation and function<br />

of the lactating mammary gland, as a luteotrophic hormone<br />

in the function of the corpus luteum and thereby as a<br />

promotor of blastocyst implantation, and an overall enhancement<br />

of the physiological functions in the tissues of the male<br />

reproductive tract (279). Nonreproductive roles of PRL include<br />

an involvement in osmoregulation; promotion of<br />

growth, development, and differentiation in several tissues;<br />

enhancement of metabolic activities in the brain, liver, pancreas,<br />

and adrenals; and various actions in immunoregulation<br />

(279).<br />

It has long been known that estradiol is a critical hormone<br />

in the regulation of PRL synthesis and secretion from the<br />

lactotrophs in the anterior pituitary. Estradiol has also been<br />

shown to stimulate lactotroph cell growth (reviewed in Ref.<br />

281) and has been implicated as a possible factor in the<br />

promotion of PRL-secreting tumors in humans (reviewed in<br />

Ref. 339). Furthermore, the lactotrophs of several species<br />

have been shown to possess significant levels of ER�,<br />

strongly suggesting that the actions of estradiol are receptormediated<br />

(reviewed in Ref. 339). As discussed above, recent<br />

descriptions have indicated the presence of ER� in the lactotrophs<br />

of the rat anterior pituitary, although contrasting<br />

reports exist, possibly due to strain variations. Whereas Wilson<br />

et al. (98) describe the presence of only ER� in lactotrophs,<br />

Mitchner et al. (99) report variable levels of ER�<br />

mRNA throughout the various cell types of the rat anterior<br />

pituitary. Shupnik et al. (100) have also reported the detection<br />

of ER� transcripts in human PRL-secreting tumors. Once<br />

again, we find only low to undetectable levels of ER� mRNA<br />

in the pituitary of the adult mouse, including the �ERKO<br />

(93).<br />

The upstream regulatory sequences of the rat PRL gene<br />

have been found to possess an estrogen-responsive element<br />

that binds ER� and functions synergistically with the pituitary-specific<br />

factor, Pit-1, to promote expression (281, 408,<br />

409). The required function of the ER� in the positive regulation<br />

of the PRL gene is nicely illustrated in the �ERKO<br />

mouse. The �ERKO females exhibit a 20-fold decrease in PRL<br />

mRNA levels in the anterior pituitary, whereas the �ERKO<br />

males exhibit a 10-fold decrease when each is compared with<br />

sex-matched wild-type controls (282). Although not as drastic,<br />

this reduction in the expression of the PRL gene is mirrored<br />

in the serum levels of the hormone in the �ERKO<br />

female, which possess an approximate 5-fold reduction in<br />

serum PRL (Table 2). Therefore, given the plethora of roles<br />

in which PRL is involved, it is likely that several of the<br />

phenotypes observed in the �ERKO mice may be due to a


June, 1999 ESTROGEN RECEPTOR NULL MICE 395<br />

direct loss of or simply enhanced by the concurrent decrease<br />

in PRL signaling.<br />

Interestingly, the extremely low levels of PRL mRNA in<br />

the anterior pituitary of the �ERKO female are even significantly<br />

less than that observed 14 days after ovariectomy in<br />

the wild-type (282). Therefore, the loss of ER� during development<br />

and differentiation of the lactotrophs in the anterior<br />

pituitary has resulted in a phenotype that is more<br />

severe than that induced by postpubertal ovariectomy, possibly<br />

due to a decrease in lactotroph cell number. It has been<br />

proposed that the lactotroph and somatotroph cell types of<br />

the adult anterior pituitary may be derived from a common<br />

cell that expresses both the genes for GH and PRL during<br />

development (reviewed in Ref. 410). The factors that may be<br />

involved in the terminal differentiation of this stem cell into<br />

a distinct cell type secreting only one of the respective hormones<br />

remain elusive. Because the appearance of the ER and<br />

the ontogeny of PRL expression appear to coincide in the<br />

developing pituitary, estrogen action has been proposed as<br />

a possible factor (410–412). However, a defect in the cell<br />

lineage of the lactotrophs that may be expected due to a loss<br />

of ER� action was not apparent in the �ERKO, as immunostaining<br />

for both PRL and GH localized expression of the<br />

genes to distinct cell types (282).<br />

Furthermore, estrogen has also been shown to stimulate<br />

proliferation of the lactotrophs and PRL-secreting cell lines<br />

(reviewed in Ref. 339). Therefore, since the marked difference<br />

in PRL mRNA levels observed between the �ERKO female<br />

and the ovariectomized wild-type is not apparently due to a<br />

defect in the differentiation of the lactotrophs, it may possibly<br />

be due to a decreased number of lactotrophs in the anterior<br />

pituitary of the �ERKO. Scully et al. (282) provided evidence<br />

against this hypothesis, by once again employing immunohistochemical<br />

staining to illustrate only a modest decrease in<br />

lactotroph cells in the anterior pituitary of the �ERKO mice.<br />

Therefore, ER� action does not appear to be required for<br />

either differentiation or proliferation of the lactotrophs in the<br />

mouse anterior pituitary. However, a recent report by Chun<br />

et al. (413) has illustrated a distinct contrast in the level of<br />

occupied ER required to elicit proliferation and that required<br />

for PRL synthesis in PR1 cells, a PRL-secreting cell line.<br />

Whereas approximately 50% of the cellular pool of ER was<br />

required to be complexed with estradiol for half-maximal<br />

stimulation of the PRL gene, only 0.1% was required to<br />

induce cellular proliferation (413). These results suggest that<br />

the mechanisms required for estrogen-induced lactotroph<br />

proliferation are hypersensitive in this cell line compared<br />

with the mechanisms involved in regulation of the PRL gene<br />

(413). Therefore, it is possible that the small amount of the<br />

active ER� splicing variant known to be present in the<br />

�ERKO (see Section II.C.) has allowed for sufficient estrogen<br />

signaling and lactotroph proliferation during develpment,<br />

resulting in the apparent lack of a somewhat expected phenotype<br />

of decreased lactotroph cell number in the pituitary<br />

of �ERKO mice.<br />

B. Behavior<br />

There are obvious effects of the gonadal steroids on sexual<br />

behavior in vertebrates; however, a more defined knowledge<br />

of these actions has become evident from a series of classical<br />

experimental schemes. These laboratory studies often relied<br />

on perinatal castration and/or developmental exposure to<br />

exogenous steroids followed by studies of the activational<br />

abilities of the different steroids during adulthood. The majority<br />

of such investigations have been carried out in the rat,<br />

but similar results have been described in other species (345,<br />

414). Breifly, studies on sexual behavior in the rat have<br />

shown that 1) castration on the day of birth results in a<br />

feminized adult male that exhibits a female pattern of behavioral<br />

responses when treated with estradiol and progesterone,<br />

and 2) neonatal testosterone or estradiol treatment of<br />

a female results in a masculinized adult that exhibits a malelike<br />

pattern of behaviors and is refractory to estradiol and<br />

progesterone (131, 337). The culmination of the data collected<br />

from such experimental schemes has led to the conclusion<br />

that testosterone secreted from the perinatal testes during a<br />

critical developmental window results in permanent changes<br />

in the hypothalamic nuclei of the brain that mediate male<br />

sexual behavior. However, the data indicating that developmental<br />

exposure to estradiol results in an adult phenotype<br />

that is similar to that elicited by testosterone suggest that<br />

many of the masculinizing effects of perinatal testosterone<br />

may be via local aromatization of the hormone to estradiol<br />

and subsequent activation of the ER signaling pathway (reviewed<br />

in Refs. 406 and 407). In addition, estradiol is also<br />

necessary for normal development of the female brain, although<br />

in lower amounts (131). Therefore, sex steroid-mediated<br />

sexual differentiation of the various regions of the<br />

brain that are critical to behavior relies not only on the nature<br />

of the steroid ligand, but also on the dose and timing of<br />

exposure (348).<br />

Before the availability of the �ERKO mouse, McCarthy et<br />

al. (415) employed an elaborate technique of infusing anti-<br />

ER� oligodeoxynucleotides into the neonatal rat hypothalamus<br />

to elucidate a direct role for ER� in the sexual differentiation<br />

of the female brain. This experimental scheme was<br />

based on the hypothesis that the presence of specific ER�<br />

antisense oligodeoxynucleotides in the hypothalamus would<br />

interfere with proper expression of the ER� gene during a<br />

critical period of differentiation (415). The experimental<br />

groups included neonatal rats treated with testosterone plus<br />

or minus the infusion of the ER� antisense oligodeoxynucleotides.<br />

As adults, those females infused with the ER� antisense<br />

oligodeoxynucleotides exhibited more female sexual<br />

behavior compared with those treated with androgen alone.<br />

The investigators thereby concluded that the reduced ER�<br />

expression protected the infused female rats from the masculinizing<br />

effects of testosterone exposure (415), providing<br />

strong evidence that local aromatization and subsequent estradiol<br />

activation of the ER� pathway plays a primary role<br />

in the masculinization of the rat brain. However, the experimental<br />

scheme of McCarthy et al. does not allow for a direct<br />

comparison with the �ERKO female, due to the caveats discussed<br />

(see Section II.C.1). It is important to recognize that the<br />

�ERKO are deficient in ER� throughout development,<br />

whereas McCarthy’s scheme produced a lack of ER� action<br />

that was only transient and most likely not as complete.<br />

The use of technologies to target individual genes has<br />

created numerous models available for studies in the behav-


396 COUSE AND KORACH Vol. 20, No. 3<br />

ioral sciences. A recent review by Nelson and Young (128)<br />

summarized and compared the behavioral or lack of behavioral<br />

phenotypes in a select 50 murine knockout models. Due<br />

to the lack of any grossly apparent behavioral phenotypes in<br />

the �ERKO mice, only those studies concerning the �ERKO<br />

will be discussed here.<br />

1. �ERKO female. The dependence of female sexual behavior<br />

on the synchronized fluctuations in estradiol and progesterone<br />

that occur during the ovarian cycle have been described<br />

in detail (reviewed in Ref. 416). In the rodent, circulating<br />

estrogens continue to rise as ovulation approaches, eventually<br />

leading to the gonadotropin surge that not only triggers<br />

the release of the oocyte from the ovary but also a marked<br />

increase in serum progesterone. This dramatic rise in circulating<br />

progesterone is required for an optimal display of the<br />

lordosis posture, a measurable response required for successful<br />

copulation (416). The gonadotropin surge from the<br />

hypothalamic-pituitary axis is due to the postive-feedback<br />

actions of estradiol. The development of this pathway in the<br />

rodent is unique to the female as a result of sexual differentiation<br />

of the neurons in the anteroventral periventricular<br />

nucleus of the preoptic area that serve to regulate hypothalamic<br />

function (264, 365, 405). As discussed above, feminization<br />

of the brain involves the actions of estradiol during<br />

fetal and neonatal development that may rely heavily on the<br />

dose and time of exposure. Therefore, knockout models for<br />

ER� and ER�, as well as the PR, have and will continue to<br />

serve as invaluable resources for dissecting the role of each<br />

receptor-signaling system in female sexual behavior.<br />

In general, evaluation of the aberrant sexual behaviors of<br />

the �ERKO female must consider not only the absence of ER�<br />

signaling, but also the elevated levels of serum testosterone<br />

that exist in the adult female (Table 2). Despite the presence<br />

of the hormones presumably required for sexual behavior,<br />

intact adult �ERKO females exhibit behavior that resembles<br />

that of a male in terms of parental, aggressive, and sexual<br />

activities (417, 418). When placed in the presence of a stud<br />

male, �ERKO females display a complete lack of sexual receptivity,<br />

measured as prelordotic behavior and a lordosis<br />

posture (418). In fact, intact �ERKO females were often<br />

treated as intruders and attacked by the stud male (417).<br />

However, similar studies using ovariectomized females indicate<br />

that this behavior of the stud male was most likely<br />

elicited by the significantly elevated levels of circulating testosterone<br />

in the �ERKO female (418, 426). These same studies,<br />

employing ovariectomized females coupled with steroid<br />

replacement of varied combinations, illustrated a complete<br />

resistance to estradiol (418, 419) and a minimal effect of<br />

progesterone in inducing a lordosis response in the �ERKO<br />

female (418).<br />

Although the above studies have indicated a prominent<br />

role for ER� in sexual behavior in the mouse, the precise<br />

pathways disrupted by a lack of ER� remain unclear. Not<br />

surprisingly, the PRKO female mice also exhibit a lack of<br />

normal sexual behavior and are unable to produce a lordosis<br />

posture even when treated with doses of either estradiol<br />

and/or progesterone (44). However, this same study reported<br />

an inability of estradiol alone to induce lordosis but<br />

rather a requirement for both estradiol and progesterone in<br />

wild-type females (44). This is in contrast to the capacity of<br />

estradiol to solely induce lordosis in the wild-type controls<br />

employed by Rissman et al. (419) and may be a reflection of<br />

the differences in background strain and/or experimental<br />

design employed in the two studies.<br />

The �ERKO and PRKO models nicely illustrate a requirement<br />

for both estradiol and progesterone action for a full<br />

lordotic response. As early as 1939, estrogen exposure before<br />

progesterone treatment was found to be required for a full<br />

display of sexual behavior in the rat (420). As in the uterus,<br />

the expression and induction of PR in certain regions of the<br />

brain is under estrogen regulation. Studies have indicated<br />

that estradiol elicits detectable increases in PR in the hypothalamus,<br />

strongly suggesting that this estrogen-action is<br />

required for ultimate progesterone-induced sexual behaviors<br />

(346, 421, 422). Therefore, disruption of the ER� gene may be<br />

expected to cause abnormally low levels of PR expression in<br />

those areas of the brain that mediate female sexual behavior<br />

and therefore may explain the lack of such behavior in the<br />

�ERKO mouse. However, two separate reports have described<br />

the ability of estradiol to induce increases in PR<br />

transcripts in the forebrain of the �ERKO female mouse,<br />

including regions of the preoptic area (400) (Fig. 8) arcuate<br />

nucleus, caudal ventromedial hypothalamus, and posterodorsal<br />

medial hypothalamus (423). However, the extent<br />

of estrogen-induced increases in hypothalamic PR in the<br />

�ERKO female is slightly attenuated compared with that<br />

observed in similarly treated wild-type mice (400, 423). It is<br />

possible that this observed estrogen action in the �ERKO<br />

female is either mediated by a splicing variant of the disrupted<br />

ER� gene or by ER�. Regardless, the level of estrogeninduced<br />

PR in the �ERKO female does not appear to allow<br />

for a response to progesterone that is sufficient to elicit sexual<br />

behavior. These data support the conclusion that normal<br />

expression of female sex behavior requires the sequential<br />

activation of the ER�- and PR-signaling pathways.<br />

Significant deficits in parental behavior and a greater tendency<br />

toward infanticide is also observed in �ERKO females<br />

compared with wild-type littermates (418). These phenotypes<br />

do not dramatically differ between intact and ovariectomized<br />

�ERKO females; however, levels of infanticide<br />

were reduced in tests carried out after a prolonged postgonadectomy<br />

period (65 days) (418). The �ERKO female<br />

exhibits aggressive behaviors that are significantly elevated<br />

compared with the wild-type female littermates, and in stark<br />

contrast to the dramatically reduced aggressive behavior<br />

displayed by the �ERKO male (to be discussed below) (418).<br />

Remarkably, acute treatment of ovariectomized females with<br />

estradiol resulted in the expected reduction in aggressive<br />

behavior in both the wild-type and �ERKO females (418).<br />

Previous studies have shown that whereas both testosterone<br />

and estradiol can elicit aggressive behaviors in the castrated<br />

male mouse, only testosterone is effective in the ovariectomized<br />

female mouse (424, 425). These studies, in combination<br />

with the findings in the �ERKO, indicate that differentiation<br />

as well as activation of aggressive behaviors in the<br />

female mouse are testosterone dependent. However, the preserved<br />

ability of estradiol to reduce aggression in the ovariectomized<br />

�ERKO female is puzzling and may indicate an<br />

ER�-mediated pathway.


June, 1999 ESTROGEN RECEPTOR NULL MICE 397<br />

The elevated levels of infanticide and aggressive behavior<br />

exhibited by the �ERKO females may be contributed to by<br />

the elevated levels of testosterone secreted by the acyclic<br />

ovary (Table 2). As discussed earlier, experimental evidence<br />

suggests that disruption of the ER� gene has resulted in a<br />

hypothalamic-pituitary axis with an enhanced capacity to<br />

respond to androgens in the �ERKO male. In support of this<br />

possibility, Ogawa et al. (418, 426) reported their preliminary<br />

finding of increased androgen receptor levels in the brain of<br />

the �ERKO female as early as 12 days of age.<br />

2. �ERKO male. Given the apparent role of ER�-mediated<br />

estrogen actions in the masculinization of the brain, it was<br />

expected that the �ERKO males would exhibit a female-like<br />

behavioral phenotype. Surprisingly, however, Ogawa et al.<br />

(318) observed that a lack of hypothalamic ER� during development<br />

has little effect on the sexual behavior of the intact<br />

�ERKO male in terms of mounting and sexual attraction<br />

toward wild-type females. In contrast, Wersinger et al. (427)<br />

report that tests of male sexual behavior carried out in a<br />

neutral arena, as opposed to the male’s home cage as done<br />

in the study of Ogawa et al., demonstrates that the number<br />

of mounting attempts exhibited by the �ERKO males is reduced.<br />

Interestingly, the studies of Ogawa et al. illustrated<br />

that �ERKO males, however, exhibit an almost complete lack<br />

of intromission and ejaculation, even though the number and<br />

frequency of mounts were similar to those of wild-type males<br />

(318). Furthermore, treatment of castrated males with estradiol<br />

or the nonaromatizable androgen, DHT, resulted in no<br />

differences in sexual behavior compared with the findings in<br />

the intact �ERKO males (428). These results differ from those<br />

described by Ono et al. (429) and Olsen (131) in the androgeninsensitive<br />

Tfm mouse, which exhibits no male-like sexual<br />

behavior including a lack of mounting as well as intromission<br />

and ejaculation. However, the �ERKO and Tfm males are<br />

similar in terms of exhibiting complete insensitivity to the<br />

effects of both estradiol and testosterone as behavioral activators<br />

during adulthood.<br />

The �ERKO male behavioral phenotype described above<br />

is obviously a contributing factor to the infertility that results<br />

after disruption of the ER� gene. The culmination of the<br />

studies indicate that a discrete component of sexual behavior<br />

in the male mouse, i.e., consummatory activity, is dependent<br />

on the actions of ER�, whereas testosterone or possibly ER�mediated<br />

estradiol actions may regulate motivational aspects<br />

(428). Although reports of categorical studies on sexual<br />

behavior are not available, both the �ERKO (47) and ArKO<br />

(257) male mice appear to be fertile and able to sire multiple<br />

litters, suggesting a minor role for ER� in sexual behavior.<br />

The possibility of compensatory mechanisms mediated by<br />

ER� in the �ERKO cannot be ruled out. It is interesting,<br />

however, that the ArKO males, presumably lacking physiological<br />

levels of estradiol throughout life, show no obvious<br />

deficits in sexual behavior that result in infertility (257). It<br />

might be expected, given the apparent need of ER�-mediated<br />

estrogen actions illustrated by the �ERKO, that the ArKO<br />

male would display a similar phenotype. Perhaps, exposure<br />

to maternal steroid hormones during gestation in the ArKO<br />

mouse has allowed for the proper “organization” of the<br />

neuronal circuitry regulating sexual behavior. More detailed<br />

studies may elucidate subtle behavioral phenotypes that exist<br />

in both the �ERKO and ArKO models and will help<br />

further define the precise role that estradiol and testosterone<br />

play in the regulation of sexual behavior.<br />

A dichotomy similar to that observed for the elements of<br />

male sexual behavior in the �ERKO is observed when behavioral<br />

assays for aggression and parental instincts are considered.<br />

Intact �ERKO males demonstrate a relatively normal<br />

pattern of parental behavior as measured by levels of<br />

infanticide when placed in the presence of newborn pups<br />

(428). However, despite the fact that �ERKO males possess<br />

serum testosterone levels that exceed the norm by as much<br />

as 2-fold and show no reduction in the levels of AR (430) or<br />

ER� (93, 352) in the brain, they consistently exhibit a significant<br />

deficit in all male aggression indices tested (428, 430).<br />

Therefore, as in the case of certain components of sexual<br />

behavior, ER�-mediated actions appear critical to the development<br />

and/or activation of aggressive behaviors, whereas<br />

parental instincts appear to be independent of ER� action<br />

(428).<br />

VII. Phenotypes in Peripheral Tissues<br />

The ER and the estrogen-signaling pathway have been<br />

described in several peripheral organ systems (reviewed in<br />

Ref. 431). A discussion of the phenotypes that may occur in<br />

each of these after disruption of either of the ER genes is<br />

beyond the scope of this review. Therefore, we have chosen<br />

to focus on three areas, all of which have received great<br />

attention as sites of estrogen action that are critical to human<br />

health. These are the bone, cardiovascular system, and adipogenesis.<br />

Furthermore, studies toward specifically defining<br />

a role that ER may play in mediating the actions of<br />

estrogen in each of these systems have begun to employ the<br />

�ERKO, and eventually the �ERKO.<br />

A. Skeletal system<br />

The link between the onset of osteoporosis and the decreasing<br />

estrogen levels associated with menopause has been<br />

realized since the report of Albright et al. in 1941 (432). Postmenopausal<br />

estrogen replacement therapy is currently the<br />

most commonly prescribed drug treatment in the United<br />

States (433, 434). In most patients, the increased risks associated<br />

with long-term estrogen replacement therapy, such as<br />

breast and endometrial cancer, are strongly overshadowed<br />

by the well established reduction in the risk of osteoporosis<br />

and bone fracture (433). Bone is a dynamic tissue that is<br />

constantly being resorbed to serve as a mineral source for the<br />

body and remodeled to replace this reservoir as well as<br />

maintain skeletal strength. Osteoporosis is a defined pathology<br />

characterized by a loss in bone mass and strength and<br />

is believed to be due to a disruption in the equilibrium<br />

between bone resorption and formation (435). Current evidence<br />

supports the hypothesis that excess bone resorption<br />

occurs in the postmenopausal years, acting to strip the bone<br />

of mass and further remove the foundation upon which new<br />

bone may be formed (435). Several therapies are known to<br />

reduce the postmenopausal increases in bone resorption,<br />

including the intake of calcium and vitamin D, calcitonin,


398 COUSE AND KORACH Vol. 20, No. 3<br />

bisphosphates, and estrogens (435). The obvious beneficial<br />

effects of estrogen replacement therapy have evoked an intense<br />

research effort for bone-specific estrogen agonists that<br />

lack the potentially harmful side effects in the breast and<br />

reproductive tissues. Ironically, the currently available “selective<br />

ER modulators” or SERMs, as these drugs have come<br />

to be termed, have been selected from the pool of nonsteroidal<br />

estrogen antagonists (reviewed in Refs. 8, 434, and<br />

436).<br />

For several years it was believed that the effects of estrogens<br />

on bone physiology were indirect, inferred from the<br />

inability to detect ER within bone and bone cell cultures.<br />

However, in 1988, Komm et al. (437) and Eriksen et al. (438)<br />

simultaneously reported the detection of high-affinity, competable<br />

estradiol binding, ER� mRNA, and the induction of<br />

estrogen-responsive genes in cultured rat and human osteoblast-like<br />

cells, the bone-forming cell. We have reported similar<br />

findings, including the inhibition of estradiol transactivational<br />

activity with antiestrogens, in two separate<br />

osteoblast-like cell lines from the rat (439). Oursler et al. (440)<br />

have since demonstrated ER� and estrogenic activity in cultured<br />

avian osteoclasts, the bone-resorbing cell type, including<br />

estrogen induction of the genes for c-fos and c-jun. Using<br />

in situ RT-PCR analysis, Hoyland et al. (441) demonstrated<br />

the presence of ER� mRNA in both osteoblasts and osteoclasts<br />

in bone grafts from human females. Immunocytochemical<br />

methods have also been used to demonstrate the<br />

presence of ER� in multiple bone cell lines (442). Recently,<br />

Bodine et al. (443) reported significant increases in the levels<br />

of ER� transcripts during dexamethasone-induced differentiation<br />

of rat osteoblasts in vitro. Therefore, there is adequate<br />

experimental evidence to support the presence of a direct<br />

ER�-mediated estrogen-signaling pathway in bone.<br />

The discovery of the ER� introduced renewed vigor in the<br />

search for SERMs, allowing for the greater possibility of<br />

finding a receptor-selective agonist. The distinct expression<br />

pattern of the two ERs among various tissues has further<br />

enhanced the possibility of finding tissue-specific SERMs.<br />

Several recent studies have reported the detection of ER� in<br />

bone cells. Onoe et al. (444) employed RT-PCR to demonstrate<br />

the presence of both ER� and ER� mRNA in immortalized<br />

as well as primary osteoblast cell cultures from the rat.<br />

Similar to reports of ER�, both Onoe et al. (444) and Arts et<br />

al. (95) report significant increases in ER� mRNA levels during<br />

in vitro dexamethasone-induced differentiation of osteoblasts<br />

derived from the rat and human, respectively. Therefore,<br />

the generation of mice lacking ER� or ER� will once<br />

again prove invaluable in delineating the roles of the two<br />

receptors in bone physiology.<br />

The majority of animal studies concerning the role of estrogens<br />

in bone morphology and metabolism have been carried<br />

out in the rat (reviewed in Ref. 445). Ovariectomy in the<br />

rodent results in increased bone turnover similar to that seen<br />

in postmenopausal women; however, the mechanisms of<br />

action may differ between the species (445). The effects of<br />

ovariectomy in the rat include decreases in bone mineral<br />

density, cancellous bone area, and bone strength, whereas<br />

increases are observed in radial and longitudinal growth,<br />

osteoblast and osteoclast activity, and overall rates of bone<br />

turnover (445). <strong>Estrogen</strong> replacement, including those com-<br />

pounds with mixed agonist/antagonist activity, has been<br />

shown to reverse several of the effects induced by ovariectomy<br />

(445). However, the extent and direction of the changes<br />

induced by ovariectomy, as well as the protection provided<br />

by estrogen replacement, vary depending on the bone parameter,<br />

sex, and type of bone being evaluated, e.g., femur,<br />

tibia, calvaria, or vertebrae (445). Interestingly, a study of the<br />

androgen-resistant Tfm rat describes a bone phenotype similar<br />

to a wild-type female, i.e., shorter and thinner femurs,<br />

indicating that androgen action may also be critical to longitudinal<br />

and radial bone growth in the male rat (446). However,<br />

endogenous gonadal estrogens were able to maintain<br />

a normal cancellous bone mass in the Tfm rat (446). It is<br />

noteworthy that the first description of a human case of<br />

estrogen insensitivity due to a spontaneous mutation of the<br />

ER� gene exhibits severe osteoporosis as well as significant<br />

increases in longitudinal growth of bones (see Section VIII.C)<br />

(116).<br />

Unfortunately, few studies of the effects of steroids on<br />

bone physiology have been carried out in the mouse. Analysis<br />

of femoral bone length in �ERKO mice indicates a significant<br />

decrease in length and diameter in females and a<br />

slight decrease in males, when compared with age- and sexmatched<br />

wild-type controls (447). However, measurements<br />

of bone density and mineral content indicate the opposite<br />

effect, i.e., �ERKO males exhibited significant decreases<br />

throughout the femur (448), whereas the �ERKO females<br />

demonstrate just slight and localized decreases (447). In<br />

agreement with the ovariectomized rat model, �ERKO female<br />

mice exhibit increased bone resorption-remodeling<br />

rates (448). However, the decreased femur length observed<br />

in the �ERKO is in contrast to that reported in the ovariectomized<br />

rat and the ER�-deficient human male. Interestingly,<br />

a series of studies by Migliaccio et al. (449, 450) illustrated<br />

that prenatal and neonatal exposure to the synthetic<br />

estrogen, DES, also results in significantly shorter femur<br />

lengths as well as increased cortical bone thickness and increased<br />

trabecular bone at the epiphysis of the femur during<br />

adulthood in female mice. Therefore, it appears that in the<br />

mouse, aberrant estrogen exposure during development or<br />

a hereditary loss of ER� action leads to decreased longitudinal<br />

bone growth, contrasting experimental schemes resulting<br />

in a similar phenotype.<br />

These data indicate that pathways other than ER� may<br />

mediate the negative regulatory effects of estradiol on bone<br />

growth, suggesting a possible role for ER�. Longitudinal<br />

bone growth is a poorly understood process that depends on<br />

chondrocyte activity, including proliferation, hypertrophy,<br />

and the secretion of extracellular matrix at the growth plate<br />

(445). <strong>Estrogen</strong> is thought to slow this process by reducing<br />

the recruitment, proliferation, and synthetic activity of chondrocytes,<br />

thereby resulting in a maturation of the epiphyseal<br />

plate and inhibition of further longitudinal growth (445). The<br />

detection of both ER� (451) and ER� (452) in human epiphyseal<br />

chondrocytes has recently been described. Therefore, it<br />

is possible that ER�, in the context of significantly elevated<br />

levels of estradiol, results in an inhibition of long bone<br />

growth in the �ERKO mouse. It is also possible that the<br />

significantly elevated levels of serum androgens in the<br />

�ERKO female may be playing an influential role. This may


June, 1999 ESTROGEN RECEPTOR NULL MICE 399<br />

also be true in the �ERKO male, which exhibit slightly<br />

shorter femur lengths in the context of only a 2-fold increase<br />

in serum androgens. The possibility that a loss of ER� during<br />

bone development results in abnormal genomic imprinting<br />

and increased ER� and/or AR levels as a compensatory<br />

mechanism must also be considered.<br />

B. Cardiovascular system<br />

Since the early part of this century, a remarkable genderrelated<br />

contrast in the risk of cardiovascular disease has been<br />

known. Although women generally possess a greater incidence<br />

of the multiple risk factors associated with cardiovascular<br />

disease when compared with men, e.g., obesity, diabetes,<br />

elevated blood pressure, and plasma cholesterol,<br />

epidemiological studies continue to indicate their relative<br />

risk of developing this disease is significantly lower (453,<br />

454). It is now believed that the protective factor against<br />

cardiovascular disease in females is their inherently increased<br />

exposure to estrogens (reviewed in Refs. 433 and<br />

454). The protective effects of estrogens have been documented<br />

in a number of epidemiological studies documenting<br />

the reduced rate of cardiovascular disease in postmenopausal<br />

women receiving estrogen replacement therapy (433,<br />

454). This correlation between the administration of estradiol<br />

and a reduced risk of vascular disease has been reproduced<br />

in laboratory animal studies involving several different species<br />

(454).<br />

The possible mechanisms by which estrogens reduce vascular<br />

disease remain unclear but are likely to include positive<br />

modifications of a number of physiological parameters that<br />

are believed to impinge upon the development of cardiovascular<br />

pathlogy (reviewed in Ref. 454). Most notable of<br />

these is the ability of estrogen replacement therapy to significantly<br />

lower total cholesterol, with a profound effect on<br />

levels of the more damaging low-density lipoproteins (LDLs)<br />

(454). Hypercholesterolemia is strongly associated with the<br />

development of cardiovascular disease. The ability of estrogen<br />

therapy to reduce total cholesterol levels is believed to<br />

account for a large portion of the cardiovascular protective<br />

effects of the hormone (433, 454). Lundeen et al. (455) reported<br />

that the significant decreases in serum cholesterol are<br />

specific to estrogen action in the ovariectomized rat, whereas<br />

other gonadal steroids tested had little effect. Furthermore,<br />

cotreatment with the antagonist, ICI-182,780, was shown to<br />

inhibit the cholesterol-lowering effects of estrogens, strongly<br />

indicating that this is a receptor-mediated effect (455). One<br />

possible site at which the actions of estradiol may converge<br />

with the pathways of cholesterol metabolism is via the upregulation<br />

of the apo E protein and the apolipoprotein (apo)<br />

B/E LDL receptor within and on the cell surface of hepatocytes,<br />

respectively (454). During hydrolysis of triglycerides<br />

in the circulation, the apo E protein is integrated into the apo<br />

B-LDL complexes and thereby functions as a ligand for the<br />

hepatocyte apo B/E LDL receptor (454). When the apo Econtaining<br />

LDL complex is bound by the apo B/E LDL receptor<br />

on the cell surface of hepatocytes, the complex is<br />

internalized, thereby providing for a means of clearing LDL<br />

from the blood (454). The importance of the apo E protein in<br />

maintaining serum cholesterol levels and providing protec-<br />

tion against vascular disease is evident from studies in transgenic<br />

mice that either overexpress the protein or possess a<br />

targeted disruption of the apo E gene. Transgenic mice in<br />

which an apo E gene is overexpressed are significantly protected<br />

from atherosclerosis (456), whereas the apo E knockout<br />

is highly susceptible to the vascular disease (457).<br />

Therefore, much of the cholesterol-lowering ability of estradiol<br />

is thought to be due to increased clearance of LDL<br />

from the circulation via the mechanism described above<br />

(454). Regulation of the apo E gene by estradiol has been<br />

demonstrated in the rat (458, 459). Furthermore, Srivastava<br />

et al. (460) recently reported that hepatic levels of apo E<br />

mRNA are similar in wild-type and �ERKO male littermates,<br />

although a slight decrease in serum levels of the protein was<br />

observed in the �ERKO. However, 6 days of estradiol exposure<br />

(via implantation of a 3 �g E 2/g body weight/day<br />

pellet) elicited an almost 2-fold increase in serum apo E levels<br />

in the wild-type compared with a 1.2-fold increase in the<br />

�ERKO (460). Therefore, these data provide evidence that<br />

ER�, acting at the level of translation, is required to mediate<br />

the estradiol up-regulation of apo E expression in mice (460).<br />

In addition to the favorable effects of estrogens on the lipid<br />

profile, it is now believed the steroid may also play a beneficial<br />

function directly at the level of the vasculature. The<br />

proposed mechanisms of estrogen action on the vasculature<br />

include modulations of vascular adhesion molecules, chemoattractants,<br />

vasodilators (e.g., nitric oxide), vasoconstrictors<br />

(e.g., endothelin-1), as well as possibly acting as an antioxidant<br />

(reviewed in Refs. 454 and 461). Several of the<br />

effects of estrogens, as well as other steroid hormones, on<br />

blood vessel physiology have been proposed to be nongenomic<br />

and independent of the classical nuclear activational<br />

pathway of steroid receptors (reviewed in Ref. 355).<br />

However, ER has been detected in both the endothelial and<br />

smooth muscle cells of the vasculature in several species,<br />

suggesting a role for receptor-mediated actions as well (reviewed<br />

in Refs. 431, 454, and 462). Furthermore, a recent<br />

study reported that the level of ER in atherosclerotic plaques<br />

from human coronary arteries was reduced compared with<br />

levels found in nonlesioned sections (463). Studies have indicated<br />

the presence of ER� in the vasculature as well. Analysis<br />

by ribonuclease protection assay for ER� and ER�<br />

mRNA in the mouse indicate only detectable levels of ER�<br />

transcripts in the aorta (93). However, Iafrati et al. (464) report<br />

the detection of ER� transcripts by RT-PCR in mouse aorta<br />

and blood vessels, including those of the �ERKO. The apparent<br />

contrast in ER� detection between these two reports<br />

in the �ERKO vasculature is most likely due to differences<br />

in the sensitivity of the techniques used, since ER� mRNA<br />

was detected only by the more sensitive RT-PCR method.<br />

This is likely a reflection of the low levels of this receptor<br />

compared with ER�. In the rat aorta, Petersen et al. (70)<br />

described the detection of full-length and variant ER�<br />

mRNA by RT-PCR. Recent reports also describe the detection<br />

of ER� transcripts by RT-PCR in aortic smooth muscle cells<br />

and coronary artery (465), aorta, and cardiac muscle (96)<br />

from monkey. An intriguing report of the tissue distribution<br />

for ER� and ER� mRNA in human vasculature was that of<br />

Tschugguel et al., which described the presence of ER�<br />

mRNA only in cultures from larger blood vessels, i.e., aorta


400 COUSE AND KORACH Vol. 20, No. 3<br />

and pulmonary artery, whereas ER� transcripts were predominant<br />

in cultures from the endothelium of smaller vessels,<br />

such as those from the uterus (466). Therefore, with the<br />

apparent variations that exist in the distribution of the two<br />

forms of ER in the vasculature, depending on the species and<br />

possibly even the vessel of study, the ERKO mice provide a<br />

unique tool to discern the direct role that ER� and/or ER�<br />

may play on vascular biology.<br />

One of the initial studies of the vasculature in the ERKO<br />

mice was that by Rubanyi et al. (467) in which the level of the<br />

vasodilator, nitric oxide, was characterized in the aortic rings<br />

of the �ERKO male. This study demonstrated that the male<br />

C57BL/6J mice, the background strain of the �ERKO, possess<br />

a significantly greater amount of ER in the aorta than<br />

female littermates, when quantified by radiolabeled E 2 binding<br />

(467). Furthermore, the basal levels of endothelial-derived<br />

nitric oxide were also found to be significantly higher<br />

in the male tissue compared with female, suggesting that the<br />

level of ER, rather than the amount of ligand, may be the most<br />

critical parameter in modulating endothelial nitric oxide production<br />

(467). In agreement with this hypothesis was the<br />

observation of significantly reduced basal nitric oxide levels<br />

in aortic tissue from male �ERKO mice (467). However, the<br />

levels of stimulated vasodilatation achieved after treatment<br />

with acetylcholine to induce endogenous nitric oxide release<br />

or nitroglycerin, a nitric oxide donor, did not differ between<br />

the sexes or ER� genotypes (467). Therefore, it appeared that<br />

only the ability to synthesize basal nitric oxide levels was<br />

altered by ER� gene disruption, whereas the capability of the<br />

vasculature to respond to the vasodilator effects of nitric<br />

oxide were not altered (467). Interestingly, the susceptibility<br />

to hypercholesterolemia-induced atherosclerosis in this<br />

strain of mice is in direct correlation with the level of ER<br />

detected, i.e., male C57BL/6J mice appear less likely to develop<br />

the vascular disease than female counterparts (468).<br />

Therefore, it may be speculated that the �ERKO mice possess<br />

an inherent susceptibility to vascular disease as well. These<br />

data also indicate that a contributing factor to the protective<br />

effects of estrogen action on the vasculature may be caused<br />

by a convergence of the estrogen and nitric oxide systems<br />

and may, in fact, be due to ligand-independent ER� actions.<br />

In contrast to the above description of a measurable vascular<br />

defect in the �ERKO mice, Mendelsohn et al. (464)<br />

demonstrated no apparent difference in the response between<br />

wild-type and �ERKO female mice in a carotid vascular<br />

injury model. This model involves the monitoring of<br />

endothelial and smooth muscle cell growth and proliferation<br />

that is spontaneously elicited in carotid arteries after artificial<br />

endothelial denudation (469). This response has been shown<br />

to be inhibited by estradiol, suggesting a mechanism by<br />

which estrogens may protect the vessel from atherosclerosis<br />

(462). Removal of estrogen action via ovariectomy in wildtype<br />

and �ERKO mice resulted in similar levels of endothelial<br />

and smooth muscle cell proliferation 2 weeks after carotid<br />

injury, as measured both morphometrically as well as with<br />

BrdU incorporation (464). However, daily estradiol treatments<br />

commencing 1 week before injury and continued for<br />

the 2 weeks after were able to inhibit the cellular proliferation<br />

to similar levels in both the wild-type and �ERKO animals<br />

(464). Therefore, the “protective” effect of estradiol illus-<br />

trated by this model system appears to be independent of<br />

ER� action and may possibly be mediated by ER�. Similar<br />

studies in the �ERKO mice are currently underway and will<br />

prove interesting in this regard.<br />

<strong>Estrogen</strong>s may also exert direct effects at the level of the<br />

heart and have been proposed to play an important role in<br />

cardiac hypertrophy and remodeling after myocardial infarction<br />

(462, 470). In a recent report, Grohé et al. (471) described<br />

the presence of both ER� and ER� as well as the<br />

P450 arom enzyme in cardiac myocytes cultured from neonatal<br />

rats. This same study demonstrated that androgen-induced<br />

up-regulation of ER� was evident in cardiac myocytes derived<br />

from female rats, whereas ER� levels were unaffected<br />

and remained stable in the cells from both sexes (471). Furthermore,<br />

estrogens may modulate cardiac contractility via<br />

regulation of Ca 2� channel activity. Previous studies have<br />

indicated that estrogen may reduce Ca 2� channel activity in<br />

several tissues; however, demonstrative studies in the heart<br />

usually involved superphysiological doses and therefore remain<br />

an issue of controversy (355, 472). However, Johnson et<br />

al. (472) demonstrated an increased number of L-type Ca 2�<br />

channels in �ERKO male heart, presumably due to a hereditary<br />

loss of ER�. Whole hearts from �ERKO males exhibited<br />

an approximate 46% increase in the number of L-type Ca 2�<br />

channels and a corresponding delay in cardiac repolarization<br />

when compared with those from wild-type (472). Therefore,<br />

estradiol ER�-mediated actions appear to modulate cardiac<br />

contractility via regulation of the number of calcium channels,<br />

possibly providing another function to complement the<br />

protective effects of estrogens in the cardiovascular system.<br />

C. Adipogenesis<br />

A role for gonadal steroid action in adipogenesis and adipocyte<br />

function has been realized for some time (473). This is<br />

evidenced by the well known gender differences that exist in the<br />

distribution of white adipose tissue in humans. Whereas men<br />

tend to accumulate fat stores in the thoracic and abdominal<br />

regions, accumulations in women are found in the upper portions<br />

of the arms and legs. Furthermore, obese woman that also<br />

exhibit androgen excess demonstrate a male pattern of fat distribution,<br />

termed android adiposity (474). Supporting evidence<br />

for a functional relationship between the gonadal and adipogenesis<br />

systems is also provided by the direct correlation that<br />

exists between nutrition, body mass, and fat content with the<br />

onset of menarche in young females (475). In addition, white<br />

adipose tissue serves not only as an energy reservoir, but is also<br />

a site of steroid storage and metabolism that can supplement<br />

gonadal and adrenal synthesis and thereby influence the overall<br />

endocrine milieu (292, 473).<br />

The roles of several members of the nuclear hormone<br />

receptor superfamily in adipogenesis have been intensely<br />

researched, especially the glucocorticoid receptor, peroxisome<br />

proliferator-activated receptor, retinoic acid receptor,<br />

and the thyroid receptor (reviewed in Ref. 473). There is little<br />

research data concerning any direct role that ER may play in<br />

adipocyte function and distribution, but it is well known that<br />

ovariectomy results in increased fat stores and body weight<br />

in females of some strains of rodents (197, 476, 477). This<br />

effect can be prevented with estrogen replacement as well as


June, 1999 ESTROGEN RECEPTOR NULL MICE 401<br />

reproduced with prolonged antiestrogen treatments in intact<br />

females (197, 476, 477). Interestingly, a similar increased<br />

body weight is observed in �ERKO females of the C57/BL<br />

background. Gross observations upon necropsy of adult<br />

�ERKO females indicate an obvious increase in the amount<br />

of white adipose fat in the pads of the mammary gland and<br />

those lining the lateral-ventral portions of the body cavity. As<br />

early as 4 months of age, �ERKO females exhibit an increased<br />

body weight compared with female wild-type littermates, at<br />

28.7 g (�0.91) vs. 23.3 g (�0.43), respectively. This difference<br />

in body weight between wild-type and �ERKO females increases<br />

at 8 months of age, at which time the average weight<br />

of �ERKO females exceeds that of age-matched wild-types<br />

by almost 35% (t test; P � 0.01). However, by 12 months of<br />

age, increases in the body weight of wild-type females appear<br />

to decrease the gap between the two genotypes.<br />

The fact that the increased fat stores in the �ERKO female<br />

are similar to those observed in the classical ovariectomized<br />

model indicate that a loss of ER�-mediated estrogen action<br />

may alter metabolism and adipocyte physiology. Fisher et al.<br />

(257) reported a similar phenotype in the ArKO female mice,<br />

in which the weight of the gonadal and mammary fat pads<br />

were increased by 50–80%. In both �ERKO and ArKO females,<br />

serum testosterone levels are substantially elevated<br />

but do not appear to have a lessening effect on fat pad weight<br />

(257). Furthermore, there are no reports of increased body<br />

weight in the androgen-resistant Tfm mice (478), indicating<br />

that androgen action may play a lesser role than estrogen in<br />

fat storage. It is unknown at this time whether a phenotype<br />

similar to the �ERKO and ArKO mice may also exist in the<br />

�ERKO mice. However, sexually mature �ERKO mice of<br />

both sexes exhibit no significant differences in body weight<br />

and appear to possess a relatively normal distribution of<br />

white adipose tissue in the peritoneum at the ages examined.<br />

Although the above evidence strongly supports a role for<br />

ER�-mediated estrogen action in adipocyte physiology, the<br />

mechanisms involved remain unclear. Adipose tissue has been<br />

shown to possess ER� (479, 480, 486–488) and the enzymes<br />

necessary for estradiol synthesis (292, 480). Recent studies have<br />

also indicated that the gonadal sex steroids can alter the activity<br />

of lipoprotein lipase, a critical enzyme in adipocyte growth and<br />

fat storage (474). Interestingly, Lubahn et al. (481) reported that<br />

�ERKO female mice fed a diet enriched with the phytoestrogen,<br />

genistein, exhibited a reduced body weight compared with<br />

�ERKO females fed the control diet. This may indicate a non-<br />

ER�-mediated yet estrogenic action of genistein, a naturally<br />

occurring isoflavone shown to possess hormone-like actions in<br />

mammalian cells that are devoid of ER (482). Therefore, the<br />

ER�-independent actions of genistein, such as the inhibition of<br />

protein tyrosine kinases and influence on growth factor action,<br />

complicate the interpretation of the above study in terms of<br />

defining a role for ER� in fat stores.<br />

VIII. Comparison with Human Disease and Models of<br />

Deficient <strong>Estrogen</strong> Action<br />

A. Ovarian carcinogenesis<br />

Ovarian cancer is the leading cause of death from gynecological<br />

cancers and accounts for 5% of all cancer deaths in<br />

Western countries (483). However, the etiology of ovarian<br />

carcinoma is complicated by paradoxes similar to those concerning<br />

breast cancer, i.e., although risk is significantly reduced<br />

with each pregnancy, the use of oral steroid contraceptives<br />

also appears to reduce the risk (483). Approximately<br />

80–90% of human ovarian cancers are derived from the ovarian<br />

surface epithelium (484), a portion of the ovary known to<br />

be rich in ER� expression. Although the causal factors remain<br />

unclear, evidence indicates that incessant ovulation, resulting<br />

in constant rupture and estrogen-mediated repair of the<br />

surface epithelium, increases the probability of spontaneous<br />

genetic abnormalities that may lead to tumorigenesis (483,<br />

484). The exact role that estrogen and the ER may play in the<br />

induction and promotion of ovarian carcinoma remains unclear<br />

(reviewed in Refs. 483–485). A number of immortalized<br />

cell lines have been generated and characterized from human<br />

ovarian tumors and exhibit varied levels and responses to<br />

estrogen agonists and antagonists (reviewed in Ref. 484).<br />

Brandenberger et al. (94) recently reported the detection of<br />

ER� and ER� mRNA in the human ovary, ovarian tumors,<br />

and ovarian tumor cell lines. Their findings include the description<br />

of ER� mRNA predominantly in the granulosa cells<br />

of normal ovaries and a marked reduction in the levels of<br />

ER� transcripts in ovarian carcinomas (94). In contrast, a cell<br />

line derived from a human ovarian surface epithelium and<br />

several human ovarian carcinomas were reported to express<br />

high levels of ER� mRNA (94). As mentioned previously,<br />

ovarian tumors are most often derived from the outer surface<br />

epithelium, whereas granulosa cell-derived carcinoma in humans<br />

is rare. Interestingly, we observe an approximately 40%<br />

incidence of granulosa/thecal cell tumors of the ovary in<br />

�ERKO females between the ages of 15–20 months. No such<br />

ovarian tumors have been observed in the wild-type or heterozygous<br />

littermates of the �ERKO. A similar incidence and<br />

type of spontaneous tumor is reported in transgenic mice<br />

possessing significantly elevated levels of LH caused by<br />

overexpression of the LH� subunit gene (254, 255). Therefore,<br />

hypergonadotropin stimulation appears to play a significant<br />

role in the etiology of this type of ovarian tumor in<br />

the mouse. The similarities in the incidence and type of<br />

tumor observed between the �ERKO and bLH�-CTP mouse<br />

indicate that ER� probably plays a minor role. Preliminary<br />

analysis in tumor samples for the �ERKO females indicates<br />

normal to elevated levels of ER� expression. Therefore, elevated<br />

gonadotropins and estradiol may result in chronic<br />

induction of the granulosa cells to proliferate and is the likely<br />

stimulus for the ovarian tumors observed in both transgenic<br />

models. Future investigations to determine the incidence of<br />

tumors in the �ERKO ovary will prove useful in further<br />

elucidating the etiology of this neoplasia.<br />

B. Chronic anovulation<br />

Targeted gene disruption of the ER genes has resulted in<br />

partial and complete anovulation in the �ERKO and �ERKO<br />

females, respectively. In the clinical setting, chronic anovulation<br />

is often categorized by the presence or lack of estrogen<br />

synthesis (204). Chronic anovulation in the absence of estrogen<br />

is diagnosed in women who experience little to no<br />

menstrual bleeding after progesterone withdrawal and is


402 COUSE AND KORACH Vol. 20, No. 3<br />

most often associated with hypogonadotropism resulting<br />

from impaired function of the hypothalamic GnRH neurons<br />

or disorders of the pituitary (204). Chronic anovulation in the<br />

presence of estrogen synthesis can be caused by several different<br />

functional abnormalities, such as Cushing’s syndrome,<br />

hyper/hypothyroidism, adrenal hyperplasia, and<br />

ovarian tumors (204). However, the majority of cases of<br />

anovulation in the presence of estrogen are associated with<br />

polycystic ovarian syndrome (PCOS), a heterogeneous series<br />

of clinical and endocrine abnormalities thought to be due to<br />

inappropriate steroid feedback regulation of the hypothalamic<br />

pituitary axis (reviewed in 204, 486, 487). Although the<br />

ovaries of both the �ERKO and �ERKO females synthesize<br />

estradiol, the inherent insensitivity of certain target tissues to<br />

the hormone may render effects similar to those of insufficient<br />

estrogen synthesis and therefore may offer some analogy<br />

to the human syndrome. The primary defect resulting in<br />

inefficient ovulation in the �ERKO appears to be due to<br />

defects directly within the ovarian tissue, more precisely, an<br />

inability to appropriately respond to estradiol in the granulosa<br />

cells of maturing follicles. However, until further studies<br />

are carried out, an impaired ability to produce an adequate<br />

gonadotropin surge must also be considered as a<br />

possible contributory factor to the decreased ovulatory rates<br />

observed in the �ERKO. Still, the reduced number of ovulations<br />

observed in the �ERKO even after stimulation with<br />

exogenous gonadotropin suggests the presence of a significant<br />

defect within the ovary.<br />

Although an analogy of the �ERKO ovarian phenotype<br />

with a human pathology may be obscure at this point, the<br />

ovarian phenotype of the �ERKO is strikingly similar to that<br />

of PCOS patients. The most common pathological symptoms<br />

of PCOS include anovulation, hirsutism, and the presence of<br />

bilateral enlarged, white, smooth, and sclerotic ovaries with<br />

a thickened outer capsule (204, 486, 487). Internal characterization<br />

of the ovaries of a PCOS patient most often indicates<br />

the presence of follicles at various stages of atresia, a strikingly<br />

hypertrophied thecal/stromal compartment, and the<br />

observation of rare or absent corpora lutea (204). The follicles<br />

presented usually possess a reduced number of granulosa<br />

cells and little aromatase activity (204). Biochemical findings<br />

are most often characterized by elevated serum androgen<br />

and LH levels, whereas progesterone and FSH levels remain<br />

low to normal (204, 486, 487). However, there appear to be<br />

wide variations in the extent and occurrence of each of these<br />

symptoms, illustrated by the diagnosis of PCOS in women<br />

with apparently normal LH levels and no menstrual abnormalities,<br />

indicating there is likely more than one etiological<br />

factor involved in this syndrome (204, 487).<br />

The onset of PCOS most often coincides with menarche in<br />

young females, suggesting the existence of endocrine abnormalities<br />

even before puberty (204, 486). The first clinical<br />

indications are dysfunctional and/or unpredictable uterine<br />

bleeding or oligo/amenorrhea (204, 486). This is often accompanied<br />

by a high incidence of hirsutism (male-pattern<br />

hair growth) in 70–90% of PCOS females, determined to be<br />

due to increased serum androgens (204, 486). However, the<br />

causal agents that lead to increased circulating steroids and<br />

the initiation of PCOS remain unclear, although a strong<br />

familial component is believed to be involved (204, 488).<br />

Franks et al. (489) have proposed that the primary cause of<br />

PCOS is ovarian in nature and that endocrine abnormalities<br />

are secondary. Others have provided evidence to indicate<br />

that inappropriate steroid feedback at the hypothalamic-pituitary<br />

axis, resulting in increased frequency and amplitude<br />

of LH pulses, results in hyperstimulation of an otherwise<br />

normal ovary (204, 486). The initiation of the abnormal steroid<br />

signaling to the hypothalamic-pituitary axis may stem<br />

from extragonadal conversion of elevated circulating androgens<br />

to estrone and estradiol (204, 486, 487). The source of the<br />

initial increases in androgen that may trigger PCOS remain<br />

unclear and may be ovarian or adrenal in nature (487). Regardless,<br />

extraglandular aromatase activity in the adipose<br />

tissue is believed to result in the initial acyclic increases in<br />

estrogens that lead to the ovarian syndrome (204, 486). Obesity<br />

is reported in 40% of PCOS patients and is thought to be<br />

a contributory factor in the adipose tissue-derived estrogen<br />

production (204, 486). The abnormally high levels of estradiol<br />

then feed back to the hypothalamic-pituitary axis in a positive<br />

fashion, resulting in chronically elevated LH levels (204,<br />

486). Increased secretion of LH leads to hypertrophy of the<br />

ovarian thecum, a hallmark of the PCOS ovary, and further<br />

increases in androgen production (204, 486). Therefore, once<br />

initiated, it is thought that ovarian-derived androgens, followed<br />

by extraglandular conversion to estrogens, act to selfperpetuate<br />

the syndrome. Treatment of PCOS patients with<br />

a GnRH antagonist has proven to reduce circulating androgen<br />

and subsequent estrogen levels, indicating the ovary as<br />

the primary source of the androgen (204, 486). Furthermore,<br />

women with PCOS can be induced to ovulate with administration<br />

of FSH, suggesting that the primary defect is extragonadal<br />

in nature (204).<br />

The similarities of the �ERKO ovarian phenotype and<br />

those associated with PCOS are worth noting. However, the<br />

initial stimulus of abnormal feedback to the hypothalamicpituitary<br />

axis that leads to the phenotype may be different in<br />

the human and �ERKO mouse. As described above in the<br />

human female, it is believed that extragonadal synthesis of<br />

estrogens results in acyclic positive regulation of LH secretion<br />

by the hypothalamic-pituitary axis. In contrast, the<br />

�ERKO ovary is the principal source of both androgen and<br />

estradiol, and acyclic increases in serum LH are primarily<br />

due to a lack of negative feedback at the hypothalamicpituitary<br />

axis. However, preliminary studies in the �ERKO<br />

female have indicated that the signaling pathways for positive<br />

regulation of the LH secretion may be intact and therefore<br />

may also contribute to the increased gonadotropin levels<br />

observed. Therefore, although the means by which the elevated<br />

levels of LH are achieved in human PCOS and the<br />

�ERKO female mouse may differ, some features of the ovarian<br />

phenotypes are similar. Certain morphological abnormalities<br />

in the ovaries are comparable, i.e., polycystic follicles<br />

with reduced numbers of granulosa cells, hypertrophied thecal<br />

and stromal tissue, and the absence of corpora lutea,<br />

although the �ERKO ovary does not exhibit the thickened<br />

outer capsule seen in PCOS patients. It is possible that encapsulation<br />

of the polycystic ovary in the human is an ER�mediated<br />

event and therefore not possible in the �ERKO.<br />

However, the LH-overexpressing bLH�-CTP mice, which<br />

possess a functional ER� gene as well as several character-


June, 1999 ESTROGEN RECEPTOR NULL MICE 403<br />

istics associated with PCOS, also do not demonstrate this<br />

phenotype (254, 255). Several of the endocrine parameters<br />

associated with human PCOS are also observed in the<br />

�ERKO, especially the presence of elevated serum androgens<br />

and LH. Although the observation of hirsutism caused by<br />

elevated serum androgens is not possible in mice, analogous<br />

biological manifestations of elevated androgen levels are<br />

exhibited in the �ERKO female, including masculinized preputial<br />

(clitoral) glands and a thickened dermis. Furthermore,<br />

obesity is reported in a large proportion of PCOS patients and<br />

is thought to be a contributory factor in the extragonadal<br />

conversion of androgens to estrogens. As described above<br />

(Section VII.C), adult �ERKO females significantly outweigh<br />

their age-matched wild-type littermates because of excessive<br />

white adipose tissue. Interestingly, insulin resistance and<br />

hyperinsulinemia are often reported in PCOS patients and<br />

thought to contribute to further androgen synthesis in the<br />

ovarian thecum (204, 487). Taylor and Lubahn (490) reported<br />

an abnormal glucose tolerance test in �ERKO females that<br />

could be corrected with ovariectomy, suggesting PCOS-like<br />

insulin resistance. Therefore, the similarities of the endocrine<br />

abnormalities and ovarian phenotypes in the �ERKO and<br />

bLH�-CTP mice with those of human PCOS patients provide<br />

support that the cause may be extragonadal in nature, and<br />

possibly due to elevated LH in the presence of normal FSH.<br />

Although the initial cause of the abnormally high LH levels<br />

in �ERKO females may differ from that postulated in the<br />

PCOS patient, the resulting effects on the ovary are similar<br />

and may allow the �ERKO to be a useful experimental model<br />

in future investigations of this human disease.<br />

C. ER� and aromatase deficiency<br />

The first and only reported case of estrogen insensitivity<br />

in a human is that of Smith et al. (116) in which a male (46,<br />

XY) was determined to be homozygous for a single-point<br />

mutation in exon 2 of the ER� gene that resulted in a premature<br />

stop codon. Similar reports of human estrogen deficiency<br />

were also lacking in the clinical literature until recently.<br />

However, five cases of aromatase deficiency and<br />

therefore a lack of estrogen synthesis have been reported in<br />

both human males (119, 120) and females (117, 118, 120, 491).<br />

Therefore, the existence of humans that exhibit insufficient<br />

estrogen action due to either a lack of functional receptor or<br />

hormone, along with the successful generation of the ER null<br />

mice, suggests that estrogen is not essential to embryonic and<br />

fetal development in mammals. However, like the ER null<br />

mice, a distinct syndrome of phenotypes is apparent in both<br />

sexes of humans lacking in estradiol or ER.<br />

All three reports of females homozygous for inactivating<br />

mutations in the P450 arom gene describe pseudohermaphroditism,<br />

i.e., 46XX genotype, the presence of internal female<br />

reproductive structures but ambiguous external genitalia<br />

(117, 120, 491). This phenotype is due to the lack of aromatase<br />

activity in the fetus and placental unit during gestation,<br />

leading to the accumulation of androgens, which in turn elicit<br />

masculinizing effects on the fetus (117, 120, 491). A similar<br />

phenotype of ambiguous external genitalia was described in<br />

an infant in which placental aromatase deficiency was determined,<br />

although enzyme activity in the fetus was not<br />

evaluated at the time of the report (118). In the two cases of<br />

a complete lack of aromatase, the mother also exhibited virilization<br />

during pregnancy (120, 491), whereas a traceable<br />

level of placental aromatase activity appeared to inhibit this<br />

symptom in the case of Conte et al. (117). In all three cases,<br />

development of the internal structures of the female reproductive<br />

tract did not appear to be altered by a lack of estrogen<br />

action (117, 120, 491), although a compensatory role fulfilled<br />

by maternal estrogens cannot be ruled out. Serum levels of<br />

androgen, androgen precursors, FSH, and LH were elevated<br />

in the P450 arom-deficient patients as early as infancy (117,<br />

491) and continued to be elevated as the females approached<br />

puberty (117, 120). At 12–14 yr of age, secondary development<br />

of the breasts, a pubertal growth spurt, and menarche<br />

were all absent, but virilization of the external genitalia was<br />

reported (117, 120). In all three cases, hyperstimulation of the<br />

ovary and the development of multifollicular cysts was illustrated,<br />

even as early as 2 yr of age (117, 120, 491). The<br />

ovarian pathology exhibited was similar to that observed in<br />

the �ERKO females and was compatible with a diagnosis of<br />

PCOS (120). Therefore, intraovarian estrogen action does not<br />

appear to play a role in the development of the multifollicular<br />

ovarian cysts in humans. <strong>Estrogen</strong> and progesterone replacement<br />

therapy alleviated the above phenotypes, resulting in<br />

normal gonadotropin and androgen levels, regression of the<br />

ovarian cysts, and in the two pubertal patients reported, the<br />

onset of breast development, a growth spurt, and menarche<br />

(117, 120, 491).<br />

The clinical syndromes exhibited by the two reported human<br />

male cases of aromatase deficiency (119, 120) and the<br />

single known human case of an inactivating mutation of the<br />

ER� gene (116) are strikingly similar. All three patients exhibited<br />

a normal onset of puberty and no gender-identity<br />

disorders, as well as normal external genitalia and prostate<br />

size (116, 119, 120). The patient lacking functional ER� exhibited<br />

a testicular volume and sperm density within the<br />

norm for men of his age (116). A normal testicular volume<br />

was also reported in the P450 arom-deficient patient of Morishima<br />

et al. (120). However, Carani et al. (119) reported small<br />

testis and severe oligozoospermia and infertility in the other<br />

male lacking aromatase, although the presence of similar<br />

findings in a brother with a normal P450 arom gene suggested<br />

other possible familial factors for this finding. Serum levels<br />

of androgens, FSH, and LH were all consistently elevated in<br />

the P450 arom-deficient males (119, 120), as well as estrone and<br />

estradiol in the ER�-deficient male (116). <strong>Estrogen</strong> replacement<br />

therapy resulted in the return of androgen and gonadotropin<br />

levels to the normal range in the P450 arom-deficient<br />

males. However, similar estrogen therapy induced no such<br />

changes in the ER�-deficient male (116). In fact, estrogen<br />

replacement therapy in the ER�-deficient male achieved circulating<br />

levels of the steroid that exceeded the norm by<br />

nearly 5-fold, yet no alleviation of the above pathologies or<br />

side effects were observed, such as impotence and gynecomastia<br />

(116). Alterations in the cardiovascular system have<br />

been reported recently in the ER�-deficient male, including<br />

dysfunctions in vasodilation (492) and premature coronary<br />

artery disease (493).<br />

The most overt phenotypes in all patients lacking sufficient<br />

estrogen action involved the skeleton. Females homozygous


404 COUSE AND KORACH Vol. 20, No. 3<br />

for mutations of the P450 arom gene displayed a delay in bone<br />

age, a significant decrease in bone mineral density of the<br />

lumbar spine, and absence of the pubertal growth spurt (117,<br />

491). Both the ER�-deficient male and the two males lacking<br />

aromatase were evaluated as adults and exhibited strikingly<br />

similar skeletal phenotypes. All three were characterized by<br />

tall stature (�95th percentile) with continued slow linear<br />

growth, a low upper/lower body segment ratio (�0.88, vs.<br />

0.96, average for men), unfused epiphyses, bone age of 14–15<br />

yr, and decreased bone density (116, 119, 120). Epiphyseal<br />

closure and increases in bone mineral density were observed<br />

after estrogen replacement therapy in the P450 arom-deficient<br />

patients (119, 120). However, as with the phenotypes described<br />

above in the ER�-deficient male, exogenous estrogen<br />

treatments resulted in no changes in bone physiology (116).<br />

Therefore, although it was once believed that estrogens were<br />

the major sex steroid influencing bone physiology in females<br />

and androgen fulfilled similar roles in the male, the phenotypes<br />

described above strongly indicate that estrogen action<br />

is critical to the pubertal growth spurt, bone mineralization,<br />

and epiphyseal maturation in both males and females.<br />

IX. Summary<br />

All scientific investigations begin with distinct objectives:<br />

first is the hypothesis upon which studies are undertaken to<br />

disprove, and second is the overall aim of obtaining further<br />

information, from which future and more precise hypotheses<br />

may be drawn. Studies focusing on the generation and use<br />

of gene-targeted animal models also apply these goals and<br />

may be loosely categorized into sequential phases that become<br />

apparent as the use of the model progresses. Initial<br />

studies of knockout models often focus on the plausibility of<br />

the model based on prior knowledge and whether the generation<br />

of an animal lacking the particular gene will prove<br />

lethal or not. Upon the successful generation of a knockout,<br />

confirmatory studies are undertaken to corroborate previously<br />

established hypotheses of the function of the disrupted<br />

gene product. As these studies continue, observations of<br />

unpredicted phenotypes or, more likely, the lack of a phenotype<br />

that was expected based on models put forth from<br />

past investigations are noted. Often the surprising phenotype<br />

is due to the loss of a gene product that is downstream<br />

from the functions of the disrupted gene, whereas the lack of<br />

an expected phenotype may be due to compensatory roles<br />

filled by alternate mechanisms. As the descriptive studies of<br />

the knockout continue, use of the model is often shifted to the<br />

role as a unique research reagent, to be used in studies that<br />

1) were not previously possible in a wild-type model; 2)<br />

aimed at finding related proteins or pathways whose existence<br />

or functions were previously masked; or 3) the subsequent<br />

effects of the gene disruption on related physiological<br />

and biochemical systems.<br />

The �ERKO mice continue to satisfy the confirmatory role<br />

of a knockout quite well. As summarized in Table 4, the<br />

phenotypes observed in the �ERKO due to estrogen insensitivity<br />

have definitively illustrated several roles that were<br />

previously believed to be dependent on functional ER�, including<br />

1) the proliferative and differentiative actions critical<br />

to the function of the adult female reproductive tract and<br />

mammary gland; 2) as an obligatory component in growth<br />

factor signaling in the uterus and mammary gland; 3) as the<br />

principal steroid involved in negative regulation of gonadotropin<br />

gene transcription and LH levels in the hypothalamic-pituitary<br />

axis; 4) as a positive regulator of PR expression<br />

in several tissues; 5) in the positive regulation of PRL<br />

synthesis and secretion from the pituitary; 6) as a promotional<br />

factor in oncogene-induced mammary neoplasia; and<br />

7) as a crucial component in the differentiation and activation<br />

of several behaviors in both the female and male.<br />

The list of unpredictable phenotypes in the �ERKO must<br />

begin with the observation that generation of an animal<br />

lacking a functional ER� gene was successful and produced<br />

animals of both sexes that exhibit a life span comparable to<br />

wild-type. The successful generation of �ERKO mice suggests<br />

that this receptor is also not essential to survival and<br />

was most likely not a compensatory factor in the survival of<br />

the �ERKO. In support of this is our recent successful generation<br />

of double knockout, or ��ERKO mice of both sexes.<br />

The precise defects in certain components of male reproduction,<br />

including the production of abnormal sperm and the<br />

loss of intromission and ejaculatory responses that were observed<br />

in the �ERKO, were quite surprising. In turn, certain<br />

estrogen pathways in the �ERKO female appear intact or<br />

unaffected, such as the ability of the uterus to successfully<br />

exhibit a progesterone-induced decidualization response,<br />

and the possible maintenance of an LH surge system in the<br />

hypothalamus. Furthermore, it is apparent that several of the<br />

�ERKO phenotypes may be aggravated by the downstream<br />

attenuation of progesterone and PRL action or even enhanced<br />

by increased androgen sensitivity, including those<br />

observed in the mammary gland, gonads, bone, and behavior.<br />

In addition, disruption of the ER� gene may have unmasked<br />

estrogen-signaling systems that were not easily detectable<br />

in the wild-type, such as 1) the apparent estrogen<br />

actions of the 4-OH-E 2 in the �ERKO uterus that are independent<br />

of ER� and ER�; 2) the ability of estrogen to induce<br />

increased levels of hypothalamic PR in the �ERKO female; 3)<br />

estradiol-induced potentiation of select neurons of the hippocampus;<br />

and 4) the protective effects of estrogen in the<br />

carotid vascular injury model. Finally, the concurrent descriptions<br />

of human mutations resulting in a lack of estrogen<br />

action due to a loss in ligand or functional receptor have<br />

illustrated the utility of the ER null mice as a model system<br />

to further understand the pathologies that result.<br />

Therefore, the �ERKO mice have illustrated the several<br />

ways in which data collected from a knockout model can<br />

quickly contribute to the current knowledge concerning the<br />

function of a particular gene. However, it is worth noting the<br />

distinct differences in thought that occurred during the generation<br />

of the �ERKO and �ERKO models. At the time the<br />

work was initiated to generate the �ERKO mice, much was<br />

known about the many roles of estrogen and the ER; therefore,<br />

several educated predictions of the possible phenotypes<br />

were possible and have since been confirmed, rejected, or<br />

reevaluated. However, the conception of the �ERKO mice<br />

occurred only 2 yr after the discovery of the ER�. Because<br />

little was known about the function and role that the ER�<br />

might play, it was difficult to make sound predictions. There-


June, 1999 ESTROGEN RECEPTOR NULL MICE 405<br />

TABLE 4. Summary of reported phenotypes in the �ERKO and �ERKO mice<br />

ERKO model Description of phenotype Ref.<br />

Lethality of mutation<br />

ER� No 46, 116<br />

ER�<br />

Fertility<br />

No 47<br />

�ERKO Both sexes are infertile 46<br />

�ERKO<br />

General<br />

Female is subfertile (reduced litter size); male is fertile 47<br />

�ERKO Exhibit normal expression of the ER� gene 93, 121, 352<br />

Female reproductive tract<br />

�ERKO Tract undergoes normal pre- and neonatal development but is insensitive to estradiol, DES, and hydroxy 46, 123, 153<br />

tamoxifen during adulthood<br />

Presence of non-ER� or -ER� receptor-mediated estrogenic pathway for 4-OH-estradiol and methoxychlor 173, 174<br />

Loss of mitogenic actions of EGF 170<br />

Responsive to mitogenic actions of androgens 154<br />

Responsive to progesterone and able to undergo artificially induced decidualization 183<br />

Ovaries undergo normal pre- and neonatal development, but are anovulatory during adulthood, exhibit<br />

multiple hemorrhagic cysts, and no corpora lutea<br />

46, 142<br />

30–40% incidence of ovarian tumors by 18 months of age Herein<br />

�ERKO Tract undergoes normal pre- and neonatal development and appears sensitive to ovarian estrogen cycling 47<br />

during adulthood<br />

Ovaries undergo normal pre- and neonatal development, but do not exhibit normal frequency of spontaneous<br />

ovulations during adulthood, exhibit a severely attenuated response to superovulation treatment with<br />

reduced numbers of oocytes and multiple trapped preovulatory follicles<br />

Mammary gland<br />

�ERKO Undergoes normal prenatal development but is insensitive to estrogen-induced development during puberty<br />

and adulthood<br />

269<br />

Responsive to exogenous progesterone and prolactin In preparation<br />

Susceptible to proto-oncogene (Wnt-1) induced ductal hyperplasia and lobuloalveolar adenocarcinoma but<br />

tumors exhibit a delayed growth rate compared to those in wild-type<br />

297<br />

�ERKO Undergoes normal prenatal and pubertal development, virgin gland is grossly indistinguishable from that of 47<br />

age-matched wild-type<br />

Undergoes normal differentiation and lactation during pregnancy and motherhood<br />

Male reproductive tract<br />

47<br />

�ERKO Tract undergoes normal pre- and neonatal development 315, 316<br />

Age-related phenotype of attenuated fluid resorption in efferent ducts leads to dilation of Rete testis, 315, 327<br />

atrophy of the seminiferous epithelium, and decreasing sperm counts<br />

Disrupted sperm function illustrated by an inability to fertilize 315<br />

Age-related decreases in testis weight Herein<br />

Age-related increases in seminal vesicle weight Herein<br />

�ERKO Undergoes normal pre- and neonatal development with no apparent defects in spermatogenesis that impede 47<br />

fertility<br />

Neuroendocrine system: females<br />

�ERKO Anterior pituitary possesses all of the expected cell types but exhibits elevated transcripts for the<br />

gonadotropin subunits (�-gonadotropin subunit, LH-�, FSH-�)<br />

282<br />

Elevated serum levels of estradiol, testosterone, and LH, but normal serum levels of progesterone and FSH 123, 252<br />

Normal lactotroph differentiation and number in the anterior pituitary, but exhibits significant deficits in<br />

transcription of the PRL gene and serum PRL levels<br />

282, Herein<br />

Medial preoptic region of the hypothalamus exhibits elevated levels of PR transcripts which reduce with 400, 423<br />

ovariectomy and return with estradiol treatment<br />

Rapid actions of estradiol on hippocampal neurons are preserved 364<br />

�ERKO Normal serum levels of estradiol Herein<br />

Neuroendocrine system: males<br />

�ERKO Anterior pituitary possesses all the expected cell types but exhibit elevated LH-� transcripts 282, 317<br />

Elevated serum levels of estradiol, testosterone, and LH, but normal serum levels of progesterone and FSH 317, Herein<br />

Behavior: females<br />

�ERKO Exhibit a lack of estradiol and progesterone-induced sexual behavior, increased aggression, and infanticide 417, 418, 419<br />

�ERKO Exhibit no defects in sexual behavior that impede fertility 47<br />

Behavior: males<br />

�ERKO Exhibit normal mounting and attraction toward wild-type females but a complete lack of intromission and 318, 427, 428<br />

ejaculation; display reduced aggression<br />

�ERKO Exhibit no defects in sexual behavior that impede fertility 47<br />

Cardiovascular<br />

�ERKO Exhibit reduced estradiol-induced angiogenesis and reduced basal levels of vascular nitric oxide 467<br />

Exhibit wild-type response to estradiol in the carotid artery injury model 464<br />

Exhibit increased expression of L-type Ca 2� channels 472<br />

Exhibit reduced response to estrogen-induced increases in serum apolipoprotein E 460<br />

Other<br />

�ERKO Both sexes exhibit growth arrest of longitudinal bones 447, 448<br />

Impaired glucose tolerance 490<br />

No apparent defect on B lymphopoiesis 507<br />

47


406 COUSE AND KORACH Vol. 20, No. 3<br />

fore, the �ERKO mice have played a significant role in confirming<br />

many of the roles thought to be fulfilled by estrogen,<br />

whereas the �ERKO mice have and will continue to provide<br />

primary insight into the functions of the ER�. The generation<br />

of both models, as well as a forthcoming description of the<br />

��ERKO, will prove invaluable in elucidating the precise<br />

roles fulfilled by each ER, as well as any possible cooperative<br />

roles the two receptors might play within the same tissue or<br />

even within the same cell.<br />

Acknowledgments<br />

The authors are grateful to several individuals for their dedication,<br />

effort, and insight that has made the work described above possible: first<br />

and foremost, our original collaborators in the generation of the ERKO<br />

mice, Dr. Oliver Smithies for his efforts leading to the generation of both<br />

ERKO models; Dr. Dennis Lubahn for the �ERKO; and Drs. John Krege,<br />

Jeff Hodges, and Jan-Åke Gustafsson and laboratory for the �ERKO. The<br />

authors are exceptionally indebted to Sylvia Curtis, Todd Washburn, Dr.<br />

Jonathan Lindzey, Dr. Wayne Bocchinfuso, Mariana Yates, Linwood<br />

Koonce, James Clarke, Page Myers, Dr. Motohiko Taki, and Dr. Sean<br />

Kimbro for their efforts and dedication. We would also like to thank our<br />

several collaborators, Drs. Ralph Cooper, Richard DiAugustine, E. Mitch<br />

Eddy, Thomas Golding, Paul Kincaide, Diane Klotz, Michael Mendolsohn,<br />

Robert Moss, Jeffrey Moyer, Sonoko Ogawa, Donald Pfaff, Gabor<br />

Rubanyi, David Schomberg, Allen Silverstone, and Harold Varmus. We<br />

would also like to acknowledge Drs. Paul Shughrue and Istvan Merchenthaler<br />

for the contribution of Fig. 8.<br />

References<br />

1. Jensen EV, DeSombre ER 1973 <strong>Estrogen</strong>-receptor interaction. Science<br />

182:126–134<br />

2. Jensen EV, Jacobson HI 1962 Basic guides to the mechanism of<br />

estrogen action. Recent Prog Horm Res 18:387–414<br />

3. Toft D, Gorski J 1966 A receptor molecule for estrogens: isolation<br />

from the rat uterus and preliminary characterization. Proc Natl<br />

Acad Sci USA 55:1574–1581<br />

4. Evans RM 1988 The steroid and thyroid hormone receptor superfamily.<br />

Science 240:889–895<br />

5. Jensen EV, Jacobson HI 1960 Fate of steroidal estrogen in target<br />

tissues. In: Pincus G, Vollmer EP (eds) Biological Activities of Steroids<br />

in Relation to Cancer. Academic Press, New York, pp 161–174<br />

6. Jensen EV 1996 Steroid hormones, receptors, and antagonists. Ann<br />

NY Acad Sci 784:1–17<br />

7. Stumpf W, Sar M 1976 Autoradiographic localization of estrogen,<br />

androgen, progestin, and glucocorticoid in “target tissues” and “nontarget”<br />

tissues. In: Pasqualini JR (ed) <strong>Receptor</strong>s and Mechanism of<br />

action of Steroid Hormones. Marcel Dekker, New York, pp 41–84<br />

8. MacGregor JI, Jordan VC 1998 Basic guide to the mechanisms of<br />

antiestrogen action. Pharmacol Rev 50:151–196<br />

9. Greene GL, Closs LE, Fleming H, DeSombre ER, Jensen EV 1977<br />

Antibodies to estrogen receptor: immunochemical similarity of<br />

estrophilin from various mammalian species. Proc Natl Acad Sci<br />

USA 74:3681–3685<br />

10. Greene GL, Fitch FW, Jensen EV 1980 Monoclonal antibodies to<br />

estrophilin: probes for the study of estrogen receptors. Proc Natl<br />

Acad Sci USA 77:157–161<br />

11. King WJ, Greene GL 1984 Monoclonal antibodies localize oestrogen<br />

receptor in the nuclei of target cells. Nature 307:745–747<br />

12. Hollenberg SM, Weinberger C, Ong ES, Cerelli G, Oro A, Lebo<br />

R, Thompson EB, Rosenfeld MG, Evans RM 1985 Primary structure<br />

and expression of a functional human glucocorticoid receptor<br />

cDNA. Nature 318:635–641<br />

13. Walter P, Green S, Greene G, Krust A, Bornert JM, Jeltsch JM,<br />

Staub A, Jensen E, Scrace G, Waterfield M, Chambon P 1985<br />

Cloning of the human estrogen receptor cDNA. Proc Natl Acad Sci<br />

USA 82:7889–7893<br />

14. Green S, Walter P, Kumar V, Krust A, Bornert JM, Argos P,<br />

Chambon P 1986 Human oestrogen receptor cDNA: sequence,<br />

expression and homology to v-erb-A. Nature 320:134–139<br />

15. Krust A, Green S, Argos P, Kumar V, Walter P, Bornert JM,<br />

Chambon P 1986 The chicken oestrogen receptor sequence: homology<br />

with v-erbA and the human oestrogen and glucocorticoid<br />

receptors. EMBO J 5:891–897<br />

16. White R, Lees JA, Needham M, Ham J, Parker M 1987 Structural<br />

organization and expression of the mouse estrogen receptor. Mol<br />

Endocrinol 1:735–744<br />

17. Koike S, Sakai M, Muramatsu M 1987 Molecular cloning and<br />

characterization of rat estrogen receptor cDNA. Nucleic Acids Res<br />

15:2499–2513<br />

18. Weiler IJ, Lew D, Shapiro DJ 1987 The Xenopus laevis estrogen<br />

receptor: sequence homology with human and avian receptors and<br />

identification of multiple estrogen receptor messenger ribonucleic<br />

acids. Mol Endocrinol 1:355–362<br />

19. Pakdel F, Le Guellec C, Vaillant C, Le Roux MG, Valotaire Y 1989<br />

Identification and estrogen induction of two estrogen receptors<br />

(ER) messenger ribonucleic acids in the rainbow trout liver: sequence<br />

homology with other ERs. Mol Endocrinol 3:44–51<br />

20. Kumar V, Chambon P 1988 The estrogen receptor binds tightly to<br />

its responsive element as a ligand-induced homodimer. Cell 55:<br />

145–156<br />

21. Metzger D, White JH, Chambon P 1988 The human oestrogen<br />

receptor functions in yeast. Nature 334:31–36<br />

22. Lees JA, Fawell SE, Parker MG 1989 Identification of two transactivation<br />

domains in the mouse oestrogen receptor. Nucleic Acids<br />

Res 17:5477–5488<br />

23. Kumar V, Green S, Stack G, Berry M, Jin JR, Chambon P 1987<br />

Functional domains of the human estrogen receptor. Cell 51:941–951<br />

24. Metzger D, Ali S, Bornert JM, Chambon P 1995 Characterization<br />

of the amino-terminal transcriptional activation function of the<br />

human estrogen receptor in animal and yeast cells. J Biol Chem<br />

270:9535–9542<br />

25. Danielian PS, White R, Lees JA, Parker MG 1992 Identification of<br />

a conserved region required for hormone dependent transcriptional<br />

activation by steroid hormone receptors [published erratum<br />

appears in EMBO J 1992 Jun;11(6):2366]. EMBO J 11:1025–1033<br />

26. Keightley MC 1998 Steroid receptor isoforms: exception or rule?<br />

Mol Cell Endocrinol 137:1–5<br />

27. Murphy LC, Dotzlaw H, Leygue E, Douglas D, Coutts A, Watson<br />

PH 1997 <strong>Estrogen</strong> receptor variants and mutations. J Steroid Biochem<br />

Mol Biol 62:363–372<br />

28. Miksicek RJ 1994 Steroid receptor variants and their potential role<br />

in cancer. Semin Cancer Biol 5:369–379<br />

29. Friend KE, Ang LW, Shupnik MA 1995 <strong>Estrogen</strong> regulates the<br />

expression of several different estrogen receptor mRNA isoforms<br />

in rat pituitary. Proc Natl Acad Sci USA 92:4367–4371<br />

30. Mangelsdorf DJ, Thummel C, Beato M, Herrlich P, Schutz G,<br />

Umesono K, Blumberg B, Kastner P, Mark M, Chambon P, Evans<br />

RM 1995 The nuclear receptor superfamily: the second decade. Cell<br />

83:835–839<br />

31. Shibata H, Spencer TE, Onate SA, Jenster G, Tsai SY, Tsai MJ,<br />

O’Malley BW 1997 Role of co-activators and co-repressors in the<br />

mechanism of steroid/thyroid receptor action. Recent Prog Horm<br />

Res 52:141–164<br />

32. Renaud JP, Rochel N, Ruff M, Vivat V, Chambon P, Gronemeyer<br />

H, Moras D 1995 Crystal structure of the RAR-� ligand-binding<br />

domain bound to all-trans retinoic acid. Nature 378:681–689<br />

33. Bourguet W, Ruff M, Chambon P, Gronemeyer H, Moras D 1995<br />

Crystal structure of the ligand-binding domain of the human nuclear<br />

receptor RXR-�. Nature 375:377–382<br />

34. Wagner RL, Apriletti JW, McGrath ME, West BL, Baxter JD,<br />

Fletterick RJ 1995 A structural role for hormone in the thyroid<br />

hormone receptor. Nature 378:690–697<br />

35. Brzozowski AM, Pike AC, Dauter Z, Hubbard RE, Bonn T, Engstrom<br />

O, Ohman L, Greene GL, Gustafsson JÅ, Carlquist M 1997<br />

Molecular basis of agonism and antagonism in the oestrogen receptor.<br />

Nature 389:753–758<br />

36. Shiau AK, Barstad D, Loria PM, Cheng L, Kushner PJ, Agard DA,<br />

Greene GL 1998 The structural basis of estrogen receptor/coactivator<br />

recognition and the antagonism of this interaction by tamoxifen.<br />

Cell 95:927–937


June, 1999 ESTROGEN RECEPTOR NULL MICE 407<br />

37. Patterson MN, McPhaul MJ, Hughes IA 1994 Androgen insensitivity<br />

syndrome. Baillieres Clin Endocrinol Metab 8:379–404<br />

38. Wilson JD 1992 Syndromes of androgen resistance. Biol Reprod<br />

46:168–173<br />

39. Brandon DD, Markwick AJ, Chrousos GP, Loriaux DL 1989 Glucocorticoid<br />

resistance in humans and nonhuman primates. Cancer<br />

Res 49:2203s–2213s<br />

40. Chatterjee VKK, Beck-Peccoz P 1994 Thyroid hormone resistance.<br />

Baillieres Clin Endocrinol Metab 8:267–283<br />

41. Lohnes D, Kastner P, Dierich A, Mark M, LeMeur M, Chambon<br />

P 1993 Function of retinoic acid receptor-� in the mouse. Cell<br />

73:643–658<br />

42. Mendelsohn C, Lohnes D, Decimo D, Lufkin T, LeMeur M,<br />

Chambon P, Mark M 1994 Function of the retinoic acid receptors<br />

(RARs) during development (II). Multiple abnormalities at various<br />

stages of organogenesis in RAR double mutants. Development<br />

120:2749–2771<br />

43. Lohnes D, Mark M, Mendelsohn C, Dolle P, Dierich A, Gorry P,<br />

Gansmuller A, Chambon P 1994 Function of the retinoic acid receptors<br />

(RARs) during development (I). Craniofacial and skeletal abnormalities<br />

in RAR double mutants. Development 120:2723–2748<br />

44. Lydon JP, DeMayo FJ, Funk CR, Mani SK, Hughes AR, Montgomery<br />

Jr CA, Shyamala G, Conneely OM, O’Malley BW 1995<br />

<strong>Mice</strong> lacking progesterone receptor exhibit pleiotropic reproductive<br />

abnormalities. Genes Dev 9:2266–2278<br />

45. Yoshizawa T, Handa Y, Uematsu Y, Takeda S, Sekine K, Yoshihara<br />

Y, Kawakami T, Arioka K, Sato H, Uchiyama Y, Masushige S, Fukamizu<br />

A, Matsumoto T, Kato S 1997 <strong>Mice</strong> lacking the vitamin D<br />

receptor exhibit impaired bone formation, uterine hypoplasia and<br />

growth retardation after weaning. Nat Genet 16:391–396<br />

46. Lubahn DB, Moyer JS, Golding TS, Couse JF, Korach KS, Smithies<br />

O 1993 Alteration of reproductive function but not prenatal<br />

sexual development after insertional disruption of the mouse estrogen<br />

receptor gene. Proc Natl Acad Sci USA 90:11162–11166<br />

47. Krege JH, Hodgin JB, Couse JF, Enmark E, Warner M, Mahler JF,<br />

Sar M, Korach KS, Gustafsson JÅ, Smithies O 1998 Generation<br />

and reproductive phenotypes of mice lacking estrogen receptor-�.<br />

Proc Natl Acad Sci USA 95:15677–15682<br />

48. Xu J, Qiu Y, DeMayo FJ, Tsai SY, Tsai MJ, O’Malley BW 1998<br />

Partial hormone resistance in mice with disruption of the steroid<br />

receptor coactivator-1 (SRC-1) gene. Science 279:1922–1925<br />

49. Kuiper GG, Enmark E, Pelto-Huikko M, Nilsson S, Gustafsson<br />

JÅ 1996 Cloning of a novel receptor expressed in rat prostate and<br />

ovary. Proc Natl Acad Sci USA 93:5925–5930<br />

50. Tremblay GB, Tremblay A, Copeland NG, Gilbert DJ, Jenkins<br />

NA, Labrie F, Giguere V 1997 Cloning, chromosomal localization,<br />

and functional analysis of the murine estrogen receptor-�. Mol<br />

Endocrinol 11:353–365<br />

51. Mosselman S, Polman J, Dijkema R 1996 ER�: identification and<br />

characterization of a novel human estrogen receptor. FEBS Lett<br />

392:49–53<br />

52. White R, Parker MG 1998 Molecular mechanisms of steroid hormone<br />

action. Endocr Relat Cancer 5:1–14<br />

53. Katzenellenbogen JA, Katzenellenbogen BS 1996 Nuclear hormone<br />

receptors: ligand-activated regulators of transcription and<br />

diverse cell responses. Chem Biol 3:529–536<br />

54. Katzenellenbogen JA, O’Malley BW, Katzenellenbogen BS 1996<br />

Tripartite steroid hormone receptor pharmacology: interaction<br />

with multiple effector sites as a basis for the cell- and promoterspecific<br />

action of these hormones. Mol Endocrinol 10:119–131<br />

55. Weigel NL, Zhang Y 1998 Ligand-independent activation of steroid<br />

hormone receptors. J Mol Med 76:469–479<br />

56. Menasce LP, White GR, Harrison CJ, Boyle JM 1993 Localization<br />

of the estrogen receptor locus (ESR) to chromosome 6q25.1 by FISH<br />

and a simple post-FISH banding technique. Genomics 17:263–265<br />

57. Sluyser M, Rijkers AW, de Goeij CC, Parker M, Hilkens J 1988<br />

Assignment of estradiol receptor gene to mouse chromosome 10.<br />

J Steroid Biochem 31:757–761<br />

58. Keaveney M, Klug J, Gannon F 1992 Sequence analysis of the 5�<br />

flanking region of the human estrogen receptor gene [published<br />

erratum appears in DNA Seq 1992;3(3):201]. DNA Seq 2:347–358<br />

59. Grandien K, Berkenstam A, Gustafsson JÅ 1997 The estrogen<br />

receptor gene: promoter organization and expression. Int J Biochem<br />

Cell Biol 29:1343–1369<br />

60. Fasco MJ 1998 <strong>Estrogen</strong> receptor mRNA splice variants produced<br />

from the distal and proximal promoter transcripts. Mol Cell Endocrinol<br />

138:51–59<br />

61. Sluyser M 1995 Mutations in the estrogen receptor gene. Hum<br />

Mutat 6:97–103<br />

62. Pettersson K, Grandien K, Kuiper GG, Gustafsson JÅ 1997 Mouse<br />

estrogen receptor-� forms estrogen response element-binding heterodimers<br />

with estrogen receptor-�. Mol Endocrinol 11:1486–1496<br />

63. Byers M, Kuiper GG, Gustafsson JÅ, Park-Sarge OK 1997 <strong>Estrogen</strong><br />

receptor-� mRNA expression in rat ovary: down-regulation by<br />

gonadotropins. Mol Endocrinol 11:172–182<br />

64. Enmark E, Pelto-Huikko M, Grandien K, Lagercrantz S, Lagercrantz<br />

J, Fried G, Nordenskjold M, Gustafsson JÅ 1997 Human<br />

estrogen receptor �-gene structure, chromosomal localization, and<br />

expression pattern. J Clin Endocrinol Metab 82:4258–4265<br />

65. Ogawa S, Inoue S, Watanabe T, Hiroi H, Orimo A, Hosoi T, Ouchi<br />

Y, Muramatsu M 1998 The complete primary structure of human<br />

estrogen receptor-� (hER�) and its heterodimerization with ER �<br />

in vivo and in vitro. Biochem Biophys Res Commun 243:122–126<br />

66. Bhat RA, Harnish DC, Stevis PE, Lyttle CR, Komm BS 1998 A<br />

novel human estrogen receptor-�: identification and functional<br />

analysis of additional N-terminal amino acids. J Steroid Biochem<br />

Mol Biol 67:233–240<br />

67. Leygue E, Dotzlaw H, Lu B, Glor C, Watson PH, Murphy LC 1998<br />

<strong>Estrogen</strong> receptor-�: mine is longer than yours? [letter]. J Clin<br />

Endocrinol Metab 83:3754–3755<br />

68. Moore JT, McKee DD, Slentz-Kesler K, Moore LB, Jones SA,<br />

Horne EL, Su JL, Kliewer SA, Lehmann JM, Willson TM 1998<br />

Cloning and characterization of human estrogen receptor-� isoforms.<br />

Biochem Biophys Res Commun 247:75–78<br />

69. Rosenfeld CS, Yuan X, Manikkam M, Calder MD, Garverick HA,<br />

Lubahn DB 1999 Cloning, sequencing, and localization of bovine<br />

estrogen receptor-� within the ovarian follicle. Biol Reprod 60:691–697<br />

70. Petersen DN, Tkalcevic GT, Koza-Taylor PH, Turi TG, Brown TA<br />

1998 Identification of estrogen receptor-�2, a functional variant of<br />

estrogen receptor-� expressed in normal rat tissues. Endocrinology<br />

139:1082–1092<br />

71. Chu S, Fuller PJ 1997 Identification of a splice variant of the rat<br />

estrogen receptor-� gene. Mol Cell Endocrinol 132:195–199<br />

72. Lu B, Leygue E, Dotzlaw H, Murphy LJ, Murphy LC, Watson PH<br />

1998 <strong>Estrogen</strong> receptor-� mRNA variants in human and murine<br />

tissues. Mol Cell Endocrinol 138:199–203<br />

73. Vladusic EA, Hornby AE, Guerra-Vladusic FK, Lupu R 1998 Expression<br />

of estrogen receptor-� messenger RNA variant in breast<br />

cancer. Cancer Res 58:210–214<br />

74. Parker MG 1995 Structure and function of estrogen receptors.<br />

Vitam Horm 51:267–287<br />

75. Jensen EV 1991 Overview of the nuclear receptor family. In: Parker<br />

MG (ed) Nuclear Hormone <strong>Receptor</strong>s: Molecular Mechanisms, Cellular<br />

Functions, Clinical Abnormalities. Academic Press, London,<br />

pp 1–13<br />

76. Tsai MJ, O’Malley BW 1994 Molecular mechanisms of action of<br />

steroid/thyroid receptor superfamily members. Annu Rev Biochem<br />

63:451–486<br />

77. Webb P, Lopez GN, Uht RM, Kushner PJ 1995 Tamoxifen activation<br />

of the estrogen receptor/AP-1 pathway: potential origin for<br />

the cell-specific estrogen-like effects of antiestrogens. Mol Endocrinol<br />

9:443–456<br />

78. Sukovich DA, Mukherjee R, Benfield PA 1994 A novel, cell-typespecific<br />

mechanism for estrogen receptor-mediated gene activation<br />

in the absence of an estrogen-responsive element. Mol Cell Biol<br />

14:7134–7143<br />

79. Ponglikitmongkol M, Green S, Chambon P 1988 Genomic organization<br />

of the human oestrogen receptor gene. EMBO J 7:3385–<br />

3388<br />

80. Kuiper GG, Carlsson B, Grandien K, Enmark E, Haggblad J,<br />

Nilsson S, Gustafsson JÅ 1997 Comparison of the ligand binding<br />

specificity and transcript tissue distribution of estrogen receptors-�<br />

and -�. Endocrinology 138:863–870<br />

81. Montano MM, Muller V, Trobaugh A, Katzenellenbogen BS 1995<br />

The carboxy-terminal F domain of the human estrogen receptor:


408 COUSE AND KORACH Vol. 20, No. 3<br />

role in the transcriptional activity of the receptor and the effectiveness<br />

of antiestrogens as estrogen antagonists. Mol Endocrinol<br />

9:814–825<br />

82. Peters GA, Khan SA 1999 <strong>Estrogen</strong> receptor domains E and F: role<br />

in dimerization and interaction with coactivator RIP-140. Mol Endocrinol<br />

13:286–296<br />

83. Picard D, Kumar V, Chambon P, Yamamoto KR 1990 Signal transduction<br />

by steroid hormones: nuclear localization is differentially<br />

regulated in estrogen and glucocorticoid receptors. Cell Regul<br />

1:291–299<br />

84. Ylikomi T, Bocquel MT, Berry M, Gronemeyer H, Chambon P<br />

1992 Cooperation of proto-signals for nuclear accumulation of estrogen<br />

and progesterone receptors. EMBO J 11:3681–3694<br />

85. Norris JD, Fan D, Kerner SA, McDonnell DP 1997 Identification<br />

of a third autonomous activation domain within the human estrogen<br />

receptor. Mol Endocrinol 11:747–754<br />

86. Cowley SM, Hoare S, Mosselman S, Parker MG 1997 <strong>Estrogen</strong><br />

receptors � and � form heterodimers on DNA. J Biol Chem 272:<br />

19858–19862<br />

87. Giguere V, Tremblay A, Tremblay GB 1998 <strong>Estrogen</strong> receptor-�:<br />

re-evaluation of estrogen and antiestrogen signaling. Steroids 63:<br />

335–339<br />

88. Shughrue PJ, Scrimo PJ, Merchenthaler I 1998 Evidence for the<br />

colocalization of estrogen receptor-� mRNA and estrogen receptor-�<br />

immunoreactivity in neurons of the rat forebrain. Endocrinology<br />

139:5267–5270<br />

89. Tremblay GB, Tremblay A, Labrie F, Giguere V 1998 Ligandindependent<br />

activation of the estrogen receptors-� and -� by mutations<br />

of a conserved tyrosine can be abolished by antiestrogens.<br />

Cancer Res 58:877–881<br />

90. Watanabe T, Inoue S, Ogawa S, Ishii Y, Hiroi H, Ikeda K, Orimo<br />

A, Muramatsu M 1997 Agonistic effect of tamoxifen is dependent<br />

on cell type, ERE-promoter context, and estrogen receptor subtype:<br />

functional difference between estrogen receptors-� and -�. Biochem<br />

Biophys Res Commun 236:140–145<br />

91. Paech K, Webb P, Kuiper GG, Nilsson S, Gustafsson J, Kushner<br />

PJ, Scanlan TS 1997 Differential ligand activation of estrogen receptors<br />

ER� and ER� at AP1 sites. Science 277:1508–1510<br />

92. Sun J, Meyers MJ, Fink BE, Rajendran R, Katzenellenbogen JA,<br />

Katzenellenbogen BS 1999 Novel ligands that function as selective<br />

estrogens or antiestrogens for estrogen receptor-� or estrogen receptor-�.<br />

Endocrinology 140:800–804<br />

93. Couse JF, Lindzey J, Grandien K, Gustafsson JÅ, Korach KS 1997<br />

Tissue distribution and quantitative analysis of estrogen receptor-�<br />

(ER�) and estrogen receptor-� (ER�) messenger ribonucleic acid in<br />

the wild-type and ER�- knockout mouse. Endocrinology 138:4613–<br />

4621<br />

94. Brandenberger AW, Tee MK, Jaffe RB 1998 <strong>Estrogen</strong> receptor-�<br />

(ER�) and -� (ER�) mRNAs in normal ovary, ovarian serous cystadenocarcinoma<br />

and ovarian cancer cell lines: down-regulation of<br />

ER-� in neoplastic tissues. J Clin Endocrinol Metab 83:1025–1028<br />

95. Arts J, Kuiper GG, Janssen JM, Gustafsson JÅ, Lowik CW, Pols<br />

HA, van Leeuwen JP 1997 Differential expression of estrogen receptors-�<br />

and -� mRNA during differentiation of human osteoblast<br />

SV-HFO cells. Endocrinology 138:5067–5070<br />

96. Pau CY, Pau KY, Spies HG 1998 Putative estrogen receptor-� and<br />

-� mRNA expression in male and female rhesus macaques. Mol<br />

Cell Endocrinol 146:59–68<br />

97. Shughrue PJ, Komm B, Merchenthaler I 1996 The distribution of<br />

estrogen receptor-� mRNA in the rat hypothalamus. Steroids 61:<br />

678–681<br />

98. Wilson ME, Price Jr RH, Handa RJ 1998 <strong>Estrogen</strong> receptor-� messenger<br />

ribonucleic acid expression in the pituitary gland. Endocrinology<br />

139:5151–5156<br />

99. Mitchner NA, Garlick C, Ben-Jonathan N 1998 Cellular distribution<br />

and gene regulation of estrogen receptors-� and -� in the rat<br />

pituitary gland. Endocrinology 139:3976–3983<br />

100. Shupnik MA, Pitt LK, Soh AY, Anderson A, Lopes MB, Laws Jr<br />

ER 1998 Selective expression of estrogen receptor-� and -� isoforms<br />

in human pituitary tumors. J Clin Endocrinol Metab 83:3965–3972<br />

101. Dotzlaw H, Leygue E, Watson PH, Murphy LC 1997 Expression<br />

of estrogen receptor-� in human breast tumors. J Clin Endocrinol<br />

Metab 82:2371–2374<br />

102. Dotzlaw H, Leygue E, Watson PH, Murphy LC 1999 <strong>Estrogen</strong><br />

receptor-� messenger RNA expression in human breast tumor<br />

biopsies: relationship to steroid receptor status and regulation by<br />

progestins. Cancer Res 59:529–532<br />

103. Sar M, Welsch F 1999 Differential expression of estrogen receptor-�<br />

and estrogen receptor-� in the rat ovary. Endocrinology 140:963–<br />

971<br />

104. Hiroi H, Inoue S, Watanabe T, Goto W, Orimo A, Momoeda M,<br />

Tsutsumi O, Taketani Y, Muramatsu M 1999 Differential immunolocalization<br />

of estrogen receptor-� and -� in rat ovary and uterus.<br />

J Mol Endocrinol 22:37–44<br />

105. Knauer E 1900 Die ovarientransplantation. Arch Gynakol 60:322–<br />

376<br />

106. Marshall FHA, Jolly WA 1905 Contributions to the physiology of<br />

mammalian reproduction. Phil Trans Roy Soc B 198:99–142<br />

107. Jost A 1970 General outline about reproductive physiology and its<br />

developmental background. In: Gibian H, Plotz EJ (eds) Mammalian<br />

Reproduction. Springer-Verlag, Berlin, pp 4–32<br />

108. Keeton WT, Gould JL 1986 Biological Science. WW Norton &<br />

Company, New York, p 410<br />

109. Camper SA, Saunders TL, Kendall SK, Keri RA, Seasholtz AF,<br />

Gordon DF, Birkmeier TS, Keegan CE, Karolyi IJ, Roller ML,<br />

Burrows HL, Samuelson LC 1995 Implementing transgenic and<br />

embryonic stem cell technology to study gene expression, cell-cell<br />

interactions and gene function. Biol Reprod 52:246–257<br />

110. Nishimori K, Matzuk MM 1996 Transgenic mice in the analysis of<br />

reproductive development and function. Rev Reprod 1:203–212<br />

111. George FW, Wilson JD 1988 Sex determination and differentiation.<br />

In: Knobil E, Neil JD, Ewing LL, Greenwald GS, Markert CL, Pfaff<br />

DW (eds) The Physiology of Reproduction. Raven Press, New York,<br />

pp 3–26<br />

112. Greco TL, Duello TM, Gorski J 1993 <strong>Estrogen</strong> receptors, estradiol,<br />

and diethylstilbestrol in early development: the mouse as a model<br />

for the study of estrogen receptors and estrogen sensitivity in<br />

embryonic development of male and female reproductive tracts.<br />

Endocr Rev 14:59–71<br />

113. Wu TC, Wang L, Wan YJ 1993 Detection of estrogen receptor<br />

messenger ribonucleic acid in human oocytes and cumulus-oocyte<br />

complexes using reverse transcriptase-polymerase chain reaction.<br />

Fertil Steril 59:54–59<br />

114. Wu TC, Wang L, Wan YJ 1992 Expression of estrogen receptor gene<br />

in mouse oocyte and during embryogenesis. Mol Reprod Dev 33:<br />

407–412<br />

115. Hou Q, Gorski J 1993 <strong>Estrogen</strong> receptor and progesterone receptor<br />

genes are expressed differentially in mouse embryos during preimplantation<br />

development. Proc Natl Acad Sci USA 90:9460–9464<br />

116. Smith EP, Boyd J, Frank GR, Takahashi H, Cohen RM, Specker<br />

B, Williams TC, Lubahn DB, Korach KS 1994 <strong>Estrogen</strong> resistance<br />

caused by a mutation in the estrogen-receptor gene in a man [published<br />

erratum appears in N Engl J Med 1995 Jan 12;332(2):131].<br />

N Engl J Med 331:1056–1061<br />

117. Conte FA, Grumbach MM, Ito Y, Fisher CR, Simpson ER 1994 A<br />

syndrome of female pseudohermaphroditism, hypergonadotropic<br />

hypogonadism, and multicystic ovaries associated with missense<br />

mutations in the gene encoding aromatase (P450arom). J Clin Endocrinol<br />

Metab 78:1287–1292<br />

118. Shozu M, Akasofu K, Harada T, Kubota Y 1991 A new cause of<br />

female pseudohermaphroditism: placental aromatase deficiency.<br />

J Clin Endocrinol Metab 72:560–566<br />

119. Carani C, Qin K, Simoni M, Faustini-Fustini M, Serpente S, Boyd<br />

J, Korach KS, Simpson ER 1997 Effect of testosterone and estradiol<br />

in a man with aromatase deficiency. N Engl J Med 337:91–95<br />

120. Morishima A, Grumbach MM, Simpson ER, Fisher C, Qin K 1995<br />

Aromatase deficiency in male and female siblings caused by a novel<br />

mutation and the physiological role of estrogens. J Clin Endocrinol<br />

Metab 80:3689–3698<br />

121. Rosenfeld CS, Ganjam VK, Taylor JA, Yuan X, Stiehr JR, Hardy<br />

MP, Lubahn DB 1998 Transcription and translation of estrogen<br />

receptor-� in the male reproductive tract of estrogen receptor-�<br />

knock-out and wild-type mice. Endocrinology 139:2982–2987<br />

122. Bronson SK, Smithies O 1994 Altering mice by homologous recombination<br />

using embryonic stem cells. J Biol Chem 269:27155–<br />

27158


June, 1999 ESTROGEN RECEPTOR NULL MICE 409<br />

123. Couse JF, Curtis SW, Washburn TF, Lindzey J, Golding TS,<br />

Lubahn DB, Smithies O, Korach KS 1995 Analysis of transcription<br />

and estrogen insensitivity in the female mouse after targeted disruption<br />

of the estrogen receptor gene. Mol Endocrinol 9:1441–1454<br />

124. Luetteke NC, Qiu TH, Peiffer RL, Oliver P, Smithies O, Lee DC<br />

1993 TGF� deficiency results in hair follicle and eye abnormalities<br />

in targeted and waved-1 mice. Cell 73:263–278<br />

125. Dietz HC, Valle D, Francomano CA, Kendzior Jr RJ, Pyeritz RE,<br />

Cutting GR 1993 The skipping of constitutive exons in vivo induced<br />

by nonsense mutations. Science 259:680–683<br />

126. Wakamatsu N, Kobayashi H, Miyatake T, Tsuji S 1992 A novel<br />

exon mutation in the human �-hexosaminidase � subunit gene<br />

affects 3� splice site selection. J Biol Chem 267:2406–2413<br />

127. Reed R, Maniatis T 1986 A role for exon sequences and splice-site<br />

proximity in splice-site selection. Cell 46:681–690<br />

128. Nelson RJ, Young KA 1998 Behavior in mice with targeted disruption<br />

of single genes. Neurosci Biobehav Rev 22:453–462<br />

129. Crawley JN, Paylor R 1997 A proposed test battery and constellations<br />

of specific behavioral paradigms to investigate the behavioral<br />

phenotypes of transgenic and knockout mice. Horm Behav<br />

31:197–211<br />

130. Gerlai R 1996 Gene-targeting studies of mammalian behavior: is it<br />

the mutation or the background genotype? [published erratum<br />

appears in Trends Neurosci 1996 Jul;19(7):271]. Trends Neurosci<br />

19:177–181<br />

131. Olsen KL 1992 Genetic influences on sexual behavior differentiation.<br />

In: Gerall AA, Moltz H, Ward IL (eds) Sexual Differentiation.<br />

Plenum Press, New York, pp 1–40<br />

132. Greenman DL, Dooley K, Breeden CR 1977 Strain differences in<br />

the response of the mouse to diethylstilbestrol. J Toxicol Environ<br />

Health 3:589–597<br />

133. Greenman DL, Delongchamp RR 1979 Variability of response to<br />

diethylstilbestrol: a comparison of inbred with hybrid mice. J Toxicol<br />

Environ Health 5:131–143<br />

134. Roper RJ, Griffith JS, Lyttle CR, Doerge RW, McNabb AW,<br />

Broadbent RE, Teuscher C 1999 Interacting quantitative trait loci<br />

control phenotypic variation in murine estradiol-regulated responses.<br />

Endocrinology 140:556–561<br />

135. Mani SK, Blaustein JD, O’Malley BW 1997 Progesterone receptor<br />

function from a behavioral perspective. Horm Behav 31:244–255<br />

136. Ogasawara Y, Okamoto S, Kitamura Y, Matsumoto K 1983 Proliferative<br />

pattern of uterine cells from birth to adulthood in intact,<br />

neonatally castrated, and/or adrenalectomized mice, assayed by<br />

incorporation of [ 125 I]iododeoxyuridine. Endocrinology 113:582–<br />

587<br />

137. Luo X, Ikeda Y, Parker KL 1994 A cell-specific nuclear receptor is<br />

essential for adrenal and gonadal development and sexual differentiation.<br />

Cell 77:481–490<br />

138. Parker KL 1998 The roles of steroidogenic factor 1 in endocrine<br />

development and function. Mol Cell Endocrinol 145:15–20<br />

139. Horigome T, Ogata F, Golding TS, Korach KS 1988 Estradiolstimulated<br />

proteolytic cleavage of the estrogen receptor in mouse<br />

uterus. Endocrinology 123:2540–2548<br />

140. Yamashita S, Newbold RR, McLachlan JA, Korach KS 1990 The<br />

role of the estrogen receptor in uterine epithelial proliferation and<br />

cytodifferentiation in neonatal mice. Endocrinology 127:2456–2463<br />

141. Blaustein JD 1992 Cytoplasmic estrogen receptors in rat brain:<br />

immunocytochemical evidence using three antibodies with distinct<br />

epitopes. Endocrinology 131:1336–1342<br />

142. Schomberg DW, Couse JF, Mukherykee A, Lubahn DB, Sar M,<br />

Mayo KE, Korach KS 1999 Targeted disruption of the estrogen<br />

receptor-� (ER�) gene in female mice: characterization of ovarian<br />

reponses and phenotype in the adult. Endocrinology 140:2733–2744<br />

143. Cuhna GR, Cooke PS, Bigsby R, Brody JR 1991 Ontogeny of sex<br />

steroid receptors in mammals. In: Parker MG (ed) Nuclear Hormone<br />

<strong>Receptor</strong>s: Molecular Mechanisms, Cellular Functions, Clinical<br />

Abnormalities. Academic Press, London, pp 235–268<br />

144. Yamashita S, Newbold RR, McLachlan JA, Korach KS 1989 Developmental<br />

pattern of estrogen receptor expression in female<br />

mouse genital tracts. Endocrinology 125:2888–2896<br />

145. Pasqualini JR, Sumida C 1986 Ontogeny of steroid receptors in the<br />

reproductive system. Int Rev Cytol 101:275–324<br />

146. Katzenellenbogen BS, Greger NG 1974 Ontogeny of uterine re-<br />

sponsiveness to estrogen during early development in the rat. Mol<br />

Cell Endocrinol 2:31–42<br />

147. Sheehan DM, Branham WS, Medlock KL, Olson ME, Zehr DR<br />

1981 Uterine responses to estradiol in the neonatal rat. Endocrinology<br />

109:76–82<br />

148. Kraus WL, Katzenellenbogen BS 1993 Regulation of progesterone<br />

receptor gene expression and growth in the rat uterus: modulation<br />

of estrogen actions by progesterone and sex steroid hormone antagonists.<br />

Endocrinology 132:2371–2379<br />

149. Eide A 1975 The effect of oestradiol on the DNA synthesis in<br />

neonatal mouse uterus and cervix. Cell Tissue Res 156:551–555<br />

150. Forsberg JG 1975 Late effects in the vaginal and cervical epithelia<br />

after injections of diethylstilbestrol into neonatal mice. Am J Obstet<br />

Gynecol 121:101–104<br />

151. Martin L 1980 <strong>Estrogen</strong>s, anti-estrogens and the regulation of cell<br />

proliferation in the female reproductive tract in vivo. In: McLachlan<br />

JA (ed) <strong>Estrogen</strong>s in the Environment. Elsevier North Holland, Inc,<br />

New York, pp 103–130<br />

152. Quarmby VE, Korach KS 1984 The influence of 17 �-estradiol on<br />

patterns of cell division in the uterus. Endocrinology 114:694–702<br />

153. Korach KS, Couse JF, Curtis SW, Washburn TF, Lindzey J, Kimbro<br />

KS, Eddy EM, Migliaccio S, Snedeker SM, Lubahn DB,<br />

Schomberg DW, Smith EP 1996 <strong>Estrogen</strong> receptor gene disruption:<br />

molecular characterization and experimental and clinical phenotypes.<br />

Recent Prog Horm Res 51:159–186<br />

154. Lindzey J, Curtis SW, Washburn TF, Korach KS 1996 Uterotropic<br />

effects of dihydrotestosterone in estrogen receptor knockout and<br />

wild type mice. Program of the 78th Annual Meeting of The <strong>Endocrine</strong><br />

Society, San Francisco, CA, 1996 (Abstract OR18–4), p 77<br />

155. Clark JH, Markaverich BM 1988 Actions of ovarian steroid hormones.<br />

In: Knobil E, Neil JD, Ewing LL, Greenwald GS, Markert CL,<br />

Pfaff DW (eds) The Physiology of Reproduction. Raven Press, New<br />

York, pp 675–724<br />

156. Stancel GM, Blatti P, Benson RH, Gardner RM, Kirkland JL,<br />

Ireland JS, Sidikaro J, Weil PA 1979 Hormonal control of uterine<br />

growth and regulation of responsiveness to estrogen. In: Hamilton<br />

TH, Clark JH, Sadler WA (eds) Ontogeny of <strong>Receptor</strong>s and Reproductive<br />

Hormone Action. Raven Press, New York, pp 119–132<br />

157. Korach KS 1994 Insights from the study of animals lacking functional<br />

estrogen receptor. Science 266:1524–1527<br />

158. Hardin JW, Markaverich BM, Clark JH, Padykula HA 1979 <strong>Estrogen</strong>ic<br />

stimulation of biochemical and morphological differentiation<br />

in rat uterine nuclei. In: Hamilton TH, Clark JH, Sadler WA<br />

(eds) Ontogeny of <strong>Receptor</strong>s and Reproductive Hormone Action.<br />

Raven Press, New York, pp 119–132<br />

159. Kastner P, Krust A, Turcotte B, Stropp U, Tora L, Gronemeyer H,<br />

Chambon P 1990 Two distinct estrogen-regulated promoters generate<br />

transcripts encoding the two functionally different human<br />

progesterone receptor forms A and B. EMBO J 9:1603–1614<br />

160. Kraus WL, Montano MM, Katzenellenbogen BS 1993 Cloning of<br />

the rat progesterone receptor gene 5�-region and identification of<br />

two functionally distinct promoters. Mol Endocrinol 7:1603–1616<br />

161. Liu Y, Teng CT 1992 <strong>Estrogen</strong> response module of the mouse<br />

lactoferrin gene contains overlapping chicken ovalbumin upstream<br />

promoter transcription factor and estrogen receptor-binding elements.<br />

Mol Endocrinol 6:355–364<br />

162. Tibbetts TA, Mendoza-Meneses M, O’Malley BW, Conneely OM<br />

1998 Mutual and intercompartmental regulation of estrogen receptor<br />

and progesterone receptor expression in the mouse uterus.<br />

Biol Reprod 59:1143–1152<br />

163. DiAugustine RP, Petrusz P, Bell GI, Brown CF, Korach KS,<br />

McLachlan JA, Teng CT 1988 Influence of estrogens on mouse<br />

uterine epidermal growth factor precursor protein and messenger<br />

ribonucleic acid. Endocrinology 122:2355–2363<br />

164. Huet-Hudson YM, Chakraborty C, De SK, Suzuki Y, Andrews<br />

GK, Dey SK 1990 <strong>Estrogen</strong> regulates the synthesis of epidermal<br />

growth factor in mouse uterine epithelial cells. Mol Endocrinol<br />

4:510–523<br />

165. Nelson KG, Takahashi T, Lee DC, Luetteke NC, Bossert NL, Ross<br />

K, Eitzman BE, McLachlan JA 1992 Transforming growth factor-�<br />

is a potential mediator of estrogen action in the mouse uterus.<br />

Endocrinology 131:1657–1664<br />

166. Murphy LJ, Ghahary A 1990 Uterine insulin-like growth factor-1:


410 COUSE AND KORACH Vol. 20, No. 3<br />

regulation of expression and its role in estrogen-induced uterine<br />

proliferation. Endocr Rev 11:443–453<br />

167. Hom YK, Young P, Wiesen JF, Miettinen PJ, Derynck R, Werb Z,<br />

Cunha GR 1998 Uterine and vaginal organ growth requires epidermal<br />

growth factor receptor signaling from stroma. Endocrinology<br />

139:913–921<br />

168. Nelson KG, Takahashi T, Bossert NL, Walmer DK, McLachlan JA<br />

1991 Epidermal growth factor replaces estrogen in the stimulation<br />

of female genital-tract growth and differentiation. Proc Natl Acad<br />

Sci USA 88:21–25<br />

169. Ignar-Trowbridge DM, Nelson KG, Bidwell MC, Curtis SW,<br />

Washburn TF, McLachlan JA, Korach KS 1992 Coupling of dual<br />

signaling pathways: epidermal growth factor action involves the<br />

estrogen receptor. Proc Natl Acad Sci USA 89:4658–4662<br />

170. Curtis SW, Washburn T, Sewall C, DiAugustine R, Lindzey J,<br />

Couse JF, Korach KS 1996 Physiological coupling of growth factor<br />

and steroid receptor signaling pathways: estrogen receptor knockout<br />

mice lack estrogen-like response to epidermal growth factor.<br />

Proc Natl Acad Sci USA 93:12626–12630<br />

171. Cooke PS, Buchanan DL, Young P, Setiawan T, Brody J, Korach<br />

KS, Taylor J, Lubahn DB, Cunha GR 1997 Stromal estrogen receptors<br />

mediate mitogenic effects of estradiol on uterine epithelium.<br />

Proc Natl Acad Sci USA 94:6535–6540<br />

172. Buchanan DL, Setiawan T, Lubahn DB, Taylor JA, Kurita T,<br />

Cunha GR, Cooke PS 1999 Tissue compartment-specific estrogen<br />

receptor-� participation in the mouse uterine epithelial secretory<br />

response. Endocrinology 140:484–491<br />

173. Das SK, Taylor JA, Korach KS, Paria BC, Dey SK, Lubahn DB<br />

1997 <strong>Estrogen</strong>ic responses in estrogen receptor-� deficient mice<br />

reveal a distinct estrogen signaling pathway. Proc Natl Acad Sci<br />

USA 94:12786–12791<br />

174. Ghosh D, Bagenin A, Taylor JA, Lubahn DB, Methoxychlor acts<br />

in ER�-KO mice through an ER-� independent mechanism. Program<br />

of the 80th Annual Meeting of The <strong>Endocrine</strong> Society, New<br />

Orleans, LA, 1998 (Abstract P1-569), p 236<br />

175. Teng CT, Pentecost BT, Chen YH, Newbold RR, Eddy EM,<br />

McLachlan JA 1989 Lactotransferrin gene expression in the mouse<br />

uterus and mammary gland. Endocrinology 124:992–999<br />

176. Parvizi N, Ellendorf F, Wuttke W 1981 Catacholestrogens in the<br />

brain: action on pituitary hormone secretion and catacholamine<br />

turnover. In: Fuxe K, Gustafsson JÅ, Wetterberg L (eds) Steroid<br />

Hormone Regulation of the Brain. Pergamon Press, Oxford, New<br />

York, p 107<br />

177. Fuchs E, Gliessman P, Hess DL, Pau KY, Spies HG, Wuttke W<br />

1986 Release rates of catecholamines, GABA and �-endorphin in<br />

the preoptic area and the mediobasal hypothalamus of the rhesus<br />

monkey in push-pull perfusates: correlation with blood hormone<br />

levels. Exp Brain Res 65:224–228<br />

178. Spicer LJ, Hammond JM 1989 Regulation of ovarian function by<br />

catecholestrogens: current concepts. J Steroid Biochem 33:489–501<br />

179. Paria BC, Chakraborty C, Dey SK 1990 Catechol estrogen formation<br />

in the mouse uterus and its role in implantation. Mol Cell<br />

Endocrinol 69:25–32<br />

180. Liehr JG 1997 Dual role of oestrogens as hormones and pro-carcinogens:<br />

tumour initiation by metabolic activation of oestrogens.<br />

Eur J Cancer Prev 6:3–10<br />

181. Graham JD, Clarke CL 1997 Physiological action of progesterone<br />

in target tissues. Endocr Rev 18:502–519<br />

182. Aronica SM, Kraus WL, Katzenellenbogen BS 1994 <strong>Estrogen</strong> action<br />

via the cAMP signaling pathway: stimulation of adenylate<br />

cyclase and cAMP-regulated gene transcription. Proc Natl Acad Sci<br />

USA 91:8517–8521<br />

183. Curtis SW, Clark J, Myers P, Korach KS 1999 Disruption of estrogen<br />

signaling does not prevent progesterone action in the estrogen<br />

receptor knockout mouse uterus. Proc Natl Acad Sci USA<br />

96:3646–3651<br />

184. Das SK, Chakraborty I, Paria BC, Wang XN, Plowman G, Dey SK<br />

1995 Amphiregulin is an implantation-specific and progesteroneregulated<br />

gene in the mouse uterus. Mol Endocrinol 9:691–705<br />

185. Ding YQ, Bagchi MK, Bardin CW, Bagchi IC 1995 Calcitonin gene<br />

expression in the rat uterus during pregnancy. Recent Prog Horm<br />

Res 50:373–378<br />

186. Weitlauf HM 1988 Biology of implantation. In: Knobil E, Neil JD,<br />

Ewing LL, Greenwald GS, Markert CL, Pfaff DW (eds) The Physiology<br />

of Reproduction. Raven Press, New York, pp 231–262<br />

187. Ledford BE, Rankin JC, Markwald RR, Baggett B 1976 Biochemical<br />

and morphological changes following artificially stimulated<br />

decidualization in the mouse uterus. Biol Reprod 15:529–535<br />

188. Stewart CL, Cullinan EB 1997 Preimplantation development of the<br />

mammalian embryo and its regulation by growth factors. Dev<br />

Genet 21:91–101<br />

189. Lim H, Paria BC, Das SK, Dinchuk JE, Langenbach R, Trzaskos<br />

JM, Dey SK 1997 Multiple female reproductive failures in cyclooxygenase<br />

2-deficient mice. Cell 91:197–208<br />

190. Ma L, Benson GV, Lim H, Dey SK, Maas RL 1998 Abdominal B<br />

(AbdB) Hoxa genes: regulation in adult uterus by estrogen and<br />

progesterone and repression in Mullerian duct by the synthetic<br />

estrogen diethylstilbestrol (DES). Dev Biol 197:141–154<br />

191. Cowell TP 1969 Implantation and development of mouse eggs<br />

transferred to the uteri of non-progestational mice. J Reprod Fertil<br />

19:239–245<br />

192. Galand P, Leroy F, Chretien J 1971 Effect of oestradiol on cell<br />

proliferation and histological changes in the uterus and vagina of<br />

mice. J Endocrinol 49:243–252<br />

193. Kronenberg MS, Clark JH 1985 Changes in keratin expression<br />

during the estrogen-mediated differentiation of rat vaginal epithelium.<br />

Endocrinology 117:1480–1489<br />

194. Gimenez-Conti IB, Lynch M, Roop D, Bhowmik S, Majeski P,<br />

Conti CJ 1994 Expression of keratins in mouse vaginal epithelium.<br />

Differentiation 56:143–151<br />

195. Sourla A, Luo S, Labrie C, Belanger A, Labrie F 1997 Morphological<br />

changes induced by 6-month treatment of intact and ovariectomized<br />

mice with tamoxifen and the pure antiestrogen EM-800.<br />

Endocrinology 138:5605–5607<br />

196. Luo S, Martel C, Sourla A, Gauthier S, Merand Y, Belanger A,<br />

Labrie C, Labrie F 1997 Comparative effects of 28-day treatment<br />

with the new anti-estrogen EM-800 and tamoxifen on estrogensensitive<br />

parameters in intact mice. Int J Cancer 73:381–391<br />

197. Dukes M, Chester R, Yarwood L, Wakeling AE 1994 Effects of a<br />

non-steroidal pure antioestrogen, ZM 189,154, on oestrogen target<br />

organs of the rat including bones. J Endocrinol 141:335–341<br />

198. Buchanan DL, Kurita T, Taylor JA, Lubahn DB, Cunha GR,<br />

Cooke PS 1998 Role of stromal and epithelial estrogen receptors in<br />

vaginal epithelial proliferation, stratification, and cornification. Endocrinology<br />

139:4345–4352<br />

199. Newbold RR, Bullock BC, McLachlan JA 1983 Exposure to diethylstilbestrol<br />

during pregnancy permanently alters the ovary and<br />

oviduct. Biol Reprod 28:735–744<br />

200. Newbold RR, Bullock BC, McLachlan JA 1985 Progressive proliferative<br />

changes in the oviduct of mice following developmental<br />

exposure to diethylstilbestrol. Teratogenesis Carcinog Mutagen<br />

5:473–480<br />

201. Harper MJK 1988 Gamete and zygote transport. In: Knobil E, Neil<br />

JD, Ewing LL, Greenwald GS, Markert CL, Pfaff DW (eds) The<br />

Physiology of Reproduction. Raven Press, New York, pp 103–134<br />

202. Greenwald GS 1967 Species differences in egg transport in response<br />

to exogenous estrogen. Anat Rec 157:163–72<br />

203. Baird DT 1984 The ovary. In: Austin CR, Short RV (eds) Reproduction<br />

in Mammals: Hormonal Control of Reproduction. Cambrige<br />

University Press, Cambridge, UK, pp 91–114<br />

204. Carr BR 1998 Disorders of the ovaries and female reproductive<br />

tract. In: Wilson JD, Foster DW, Kronenberg HM, Larsen PR (eds)<br />

Williams Textbook of Endocrinology. WB Saunders Company,<br />

Philadelphia, pp 751–817<br />

205. Richards JS 1980 Maturation of ovarian follicles: actions and interactions<br />

of pituitary and ovarian hormones on follicular cell differentiation.<br />

Physiol Rev 60:51–89<br />

206. Magoffin DA 1991 Regulation of differentiated functions in ovarian<br />

theca cells. Semin Reprod Endocrinol 9:321–331<br />

207. Richards JS 1994 Hormonal control of gene expression in the ovary.<br />

Endocr Rev 15:725–751<br />

208. Picton H, Briggs D, Gosden R 1998 The molecular basis of oocyte<br />

growth and development. Mol Cell Endocrinol 145:27–37<br />

209. Robker RL, Richards JS 1998 Hormonal control of the cell cycle in<br />

ovarian cells: proliferation vs. differentiation. Biol Reprod 59:476–<br />

482


June, 1999 ESTROGEN RECEPTOR NULL MICE 411<br />

210. Gore-Langton RE, Armstrong DT 1994 Follicular steroidogenesis<br />

and its control. In: Knobil E, Neil JD, Ewing LL, Greenwald GS,<br />

Markert CL, Pfaff DW (eds) The Physiology of Reproduction.<br />

Raven Press, New York, pp 571–627<br />

211. Vanderhyden BC, Cohen JN, Morley P 1993 Mouse oocytes regulate<br />

granulosa cell steroidogenesis. Endocrinology 133:423–426<br />

212. Vanderhyden BC, Tonary AM 1995 Differential regulation of progesterone<br />

and estradiol production by mouse cumulus and mural<br />

granulosa cells by A factor(s) secreted by the oocyte. Biol Reprod<br />

53:1243–1250<br />

213. Vanderhyden BC, Macdonald EA 1998 Mouse oocytes regulate<br />

granulosa cell steroidogenesis throughout follicular development.<br />

Biol Reprod 59:1296–1301<br />

214. Pencharz RI 1940 Effect of estrogens and androgens alone and in<br />

combination with chorionic gonadotropin on the ovary of the hypophysectomized<br />

rat. Science 91:554–555<br />

215. Williams PC 1940 Effect of stilbestrol on the ovaries of the hypophysectomized<br />

rat. Nature 145:388–389<br />

216. Richards JS 1975 Estradiol receptor content in rat granulosa cells<br />

during follicular development: modification by estradiol and gonadotropins.<br />

Endocrinology 97:1174–1184<br />

217. Rao MC, Midgley Jr AR, Richards JS 1978 Hormonal regulation<br />

of ovarian cellular proliferation. Cell 14:71–8<br />

218. Goldenberg RL, Vaitukaitis JL, Ross GT 1972 <strong>Estrogen</strong> and follicle<br />

stimulation hormone interactions on follicle growth in rats. Endocrinology<br />

90:1492–1498<br />

219. Bley MA, Saragueta PE, Baranao JL 1997 Concerted stimulation of<br />

rat granulosa cell deoxyribonucleic acid synthesis by sex steroids<br />

and follicle-stimulating hormone. J Steroid Biochem Mol Biol 62:<br />

11–19<br />

220. Reilly CM, Cannady WE, Mahesh VB, Stopper VS, De Sevilla<br />

LM, Mills TM 1996 Duration of estrogen exposure prior to folliclestimulating<br />

hormone stimulation is critical to granulosa cell growth<br />

and differentiation in rats. Biol Reprod 54:1336–1342<br />

221. Burghardt RC, Anderson E 1981 Hormonal modulation of gap<br />

junctions in rat ovarian follicles. Cell Tissue Res 214:181–193<br />

222. Hernandez ER, Roberts Jr CT, LeRoith D, Adashi EY 1989 Rat<br />

ovarian insulin-like growth factor I (IGF-I) gene expression is granulosa<br />

cell-selective: 5�-untranslated mRNA variant representation<br />

and hormonal regulation. Endocrinology 125:572–574<br />

223. Kaipia A, Hsueh AJ 1997 Regulation of ovarian follicle atresia.<br />

Annu Rev Physiol 59:349–363<br />

224. Hsueh AJ, Billig H, Tsafriri A 1994 Ovarian follicle atresia: a<br />

hormonally controlled apoptotic process. Endocr Rev 15:707–724<br />

225. Richards JS, Ireland JJ, Rao MC, Bernath GA, Midgley Jr AR,<br />

Reichert Jr LE 1976 Ovarian follicular development in the rat:<br />

hormone receptor regulation by estradiol, follicle stimulating hormone<br />

and luteinizing hormone. Endocrinology 99:1562–1570<br />

226. Tonetta SA, Ireland JJ 1984 Effect of cyanoketone on follicle-stimulating<br />

hormone (FSH) induction of receptors for FSH in granulosa<br />

cells of the rat. Biol Reprod 31:487–493<br />

227. Tonetta SA, diZerega GS 1989 Intragonadal regulation of follicular<br />

maturation. Endocr Rev 10:205–229<br />

228. Wang XN, Greenwald GS 1993 Hypophysectomy of the cyclic<br />

mouse. II. Effects of follicle-stimulating hormone (FSH) and luteinizing<br />

hormone on folliculogenesis, FSH and human chorionic<br />

gonadotropin receptors, and steroidogenesis. Biol Reprod 48:595–<br />

605<br />

229. Farookhi R, Desjardins J 1986 Luteinizing hormone receptor induction<br />

in dispersed granulosa cells requires estrogen. Mol Cell<br />

Endocrinol 47:13–24<br />

230. Kessel B, Liu YX, Jia XC, Hsueh AJ 1985 Autocrine role of estrogens<br />

in the augmentation of luteinizing hormone receptor formation<br />

in cultured rat granulosa cells. Biol Reprod 32:1038–1050<br />

231. Wang XN, Greenwald GS 1993 Synergistic effects of steroids with<br />

FSH on folliculogenesis, steroidogenesis and FSH- and hCG-receptors<br />

in hypophysectomized mice. J Reprod Fertil 99:403–413<br />

232. Zhuang LZ, Adashi EY, Hsueh AJ 1982 Direct enhancement of<br />

gonadotropin-stimulated ovarian estrogen biosynthesis by estrogen<br />

and clomiphene citrate. Endocrinology 110:2219–2221<br />

233. Saiduddin S, Zassenhaus HP 1977 Estradiol-17� receptors in the<br />

immature rat ovary. Steroids 29:197–213<br />

234. Kim I, Greenwald GS 1987 <strong>Estrogen</strong> receptors in ovary and uterus<br />

of immature hamster and rat: effects of estrogens. Endocrinol Jpn<br />

34:45–53<br />

235. Kawashima M, Greenwald GS 1993 Comparison of follicular estrogen<br />

receptors in rat, hamster, and pig. Biol Reprod 48:172–179<br />

236. Kim I, Greenwald GS 1987 Effect of estrogens on follicular development<br />

and ovarian and uterine estrogen receptors in the immature<br />

rabbit, guinea pig and mouse. Endocrinol Jpn 34:871–878<br />

237. Misao R, Nakanishi Y, Sun WS, Fujimoto J, Iwagaki S, Hirose R,<br />

Tamaya T 1999 Expression of oestrogen receptor-� and -� mRNA<br />

in corpus luteum of human subjects. Mol Hum Reprod 5:17–21<br />

238. Iwai T, Fujii S, Nanbu Y, Nonogaki H, Konishi I, Mori T, Okamura<br />

H 1991 Effect of human chorionic gonadotropin on the expression<br />

of progesterone receptors and estrogen receptors in rabbit<br />

ovarian granulosa cells and the uterus. Endocrinology 129:1840–<br />

1848<br />

239. Kuroda H, Terada N, Nakayama H, Matsumoto K, Kitamura Y<br />

1988 Infertility due to growth arrest of ovarian follicles in Sl/Sl t<br />

mice. Dev Biol 126:71–79<br />

240. Dong J, Albertini DF, Nishimori K, Kumar TR, Lu N, Matzuk<br />

MM 1996 Growth differentiation factor-9 is required during early<br />

ovarian folliculogenesis. Nature 383:531–535<br />

241. Kumar TR, Wang Y, Lu N, Matzuk MM 1997 Follicle stimulating<br />

hormone is required for ovarian follicle maturation but not male<br />

fertility. Nat Genet 15:201–204<br />

242. Dierich A, Sairam MR, Monaco L, Fimia GM, Gansmuller A,<br />

LeMeur M, Sassone-Corsi P 1998 Impairing follicle-stimulating<br />

hormone (FSH) signaling in vivo: targeted disruption of the FSH<br />

receptor leads to aberrant gametogenesis and hormonal imbalance.<br />

Proc Natl Acad Sci USA 95:13612–13617<br />

243. Hillier SG, Smyth CD, Whitelaw PF, Miro F, Howles CM 1995<br />

Gonadotrophin control of follicular function. Horm Res 43:216–223<br />

244. Fitzpatrick SL, Richards JS 1991 Regulation of cytochrome P450<br />

aromatase messenger ribonucleic acid and activity by steroids and<br />

gonadotropins in rat granulosa cells. Endocrinology 129:1452–1462<br />

245. Baker J, Hardy MP, Zhou J, Bondy C, Lupu F, Bellve AR, Efstratiadis<br />

A 1996 Effects of an Igf1 gene null mutation on mouse<br />

reproduction. Mol Endocrinol 10:903–918<br />

246. Sicinski P, Donaher JL, Geng Y, Parker SB, Gardner H, Park MY,<br />

Robker RL, Richards JS, McGinnis LK, Biggers JD, Eppig JJ,<br />

Bronson RT, Elledge SJ, Weinberg RA 1996 Cyclin D2 is an FSHresponsive<br />

gene involved in gonadal cell proliferation and oncogenesis.<br />

Nature 384:470–474<br />

247. Simon AM, Goodenough DA, Li E, Paul DL 1997 Female infertility<br />

in mice lacking connexin 37. Nature 385:525–529<br />

248. Matzuk MM, Kumar TR, Bradley A 1995 Different phenotypes for<br />

mice deficient in either activins or activin receptor type II. Nature<br />

374:356–360<br />

249. Matzuk MM, Dionne L, Guo Q, Kumar TR, Lebovitz RM 1998<br />

Ovarian function in superoxide dismutase 1 and 2 knockout mice.<br />

Endocrinology 139:4008–4011<br />

250. Billig H, Furuta I, Hsueh AJ 1993 <strong>Estrogen</strong>s inhibit and androgens<br />

enhance ovarian granulosa cell apoptosis. Endocrinology 133:<br />

2204–2212<br />

251. Shupnik MA 1996 Gonadal hormone feedback on pituitary gonadotropin<br />

genes. Trends Endocrinol Metab 7:272–276<br />

252. Lindzey J, Couse JF, Stoker T, Wetsel WC, Cooper R, Korach KS,<br />

Steroid regulation of gonadotrope function in female wild-type<br />

(WT) and estrogen receptor-� knockout (ERKO) mice. Program of<br />

the 80th Annual Meeting of The <strong>Endocrine</strong> Society, New Orleans,<br />

LA, 1998 (Abstract OR46-2), p 112<br />

253. Hillier SG 1994 Current concepts of the roles of follicle stimulating<br />

hormone and luteinizing hormone in folliculogenesis. Hum Reprod<br />

9:188–191<br />

254. Risma KA, Clay CM, Nett TM, Wagner T, Yun J, Nilson JH 1995<br />

Targeted overexpression of luteinizing hormone in transgenic mice<br />

leads to infertility, polycystic ovaries, and ovarian tumors. Proc<br />

Natl Acad Sci USA 92:1322–1326<br />

255. Risma KA, Hirshfield AN, Nilson JH 1997 Elevated luteinizing<br />

hormone in prepubertal transgenic mice causes hyperandrogenemia,<br />

precocious puberty, and substantial ovarian pathology. Endocrinology<br />

138:3540–3547<br />

256. Bogovich K 1992 Follicle-stimulating hormone plays a role in the


412 COUSE AND KORACH Vol. 20, No. 3<br />

induction of ovarian follicular cysts in hypophysectomized rats.<br />

Biol Reprod 47:149–161<br />

257. Fisher CR, Graves KH, Parlow AF, Simpson ER 1998 Characterization<br />

of mice deficient in aromatase (ArKO) because of targeted<br />

disruption of the cyp19 gene. Proc Natl Acad Sci USA 95:6965–6970<br />

258. Sherr CJ 1996 Cancer cell cycles. Science 274:1672–1677<br />

259. Robker RL, Richards JS 1998 Hormone-induced proliferation and<br />

differentiation of granulosa cells: a coordinated balance of the cell<br />

cycle regulators cyclin D2 and p27 Kip1 . Mol Endocrinol 12:924–940<br />

260. Natraj U, Richards JS 1993 Hormonal regulation, localization, and<br />

functional activity of the progesterone receptor in granulosa cells<br />

of rat preovulatory follicles. Endocrinology 133:761–769<br />

261. Sirois J, Richards JS 1993 Transcriptional regulation of the rat<br />

prostaglandin endoperoxide synthase 2 gene in granulosa cells.<br />

Evidence for the role of a cis-acting C/EBP� promoter element.<br />

J Biol Chem 268:21931–21938<br />

262. Richards JS, Russell DL, Robker RL, Dajee M, Alliston TN 1998<br />

Molecular mechanisms of ovulation and luteinization. Mol Cell<br />

Endocrinol 145:47–54<br />

263. Mahesh VB, Brann DW 1998 Regulation of the preovulatory gonadotropin<br />

surge by endogenous steroids. Steroids 63:616–629<br />

264. Herbison AE 1998 Multimodal influence of estrogen upon gonadotropin-releasing<br />

hormone neurons. Endocr Rev 19:302–330<br />

265. Li X, Schwartz PE, Rissman EF 1997 Distribution of estrogen receptor-�-like<br />

immunoreactivity in rat forebrain. Neuroendocrinology<br />

66:63–67<br />

266. Imagawa W, Yang J, Guzman R, Nandi S 1994 Control of mammary<br />

development. In: Knobil E, Neil JD, Ewing LL, Greenwald GS,<br />

Markert CL, Pfaff DW (eds) The Physiology of Reproduction.<br />

Raven Press, New York, pp 1033–1063<br />

267. Mori T, Nagasawa H, Bern HA 1980 Long-term effects of perinatal<br />

exposure to hormones on normal and neoplastic mammary growth<br />

in rodents: a review. J Environ Pathol Toxicol 3:191–205<br />

268. Kleinberg DL 1998 Role of IGF-I in normal mammary development.<br />

Breast Cancer Res Treat 47:201–208<br />

269. Bocchinfuso WP, Korach KS 1997 Mammary gland development<br />

and tumorigenesis in estrogen receptor knockout mice. J Mammary<br />

Gland Biol Neoplasia 2:323–334<br />

270. Daniel CW, Silberstein GB, Strickland P 1987 Direct action of 17<br />

�-estradiol on mouse mammary ducts analyzed by sustained release<br />

implants and steroid autoradiography. Cancer Res 47:6052–<br />

6057<br />

271. Silberstein GB, Van Horn K, Shyamala G, Daniel CW 1994 Essential<br />

role of endogenous estrogen in directly stimulating mammary<br />

growth demonstrated by implants containing pure antiestrogens.<br />

Endocrinology 134:84–90<br />

272. Haslam SZ, Nummy KA 1992 The ontogeny and cellular distribution<br />

of estrogen receptors in normal mouse mammary gland. J<br />

Steroid Biochem Mol Biol 42:589–595<br />

273. Dickson RB, Lippman ME 1995 Growth factors in breast cancer.<br />

Endocr Rev 16:559–589<br />

274. Nandi S, Guzman RC, Miyamoto S 1992 Hormones, cell proliferation,<br />

and mammary carcinogenesis. In: Li JJ Nandi S, Li SA (eds)<br />

Hormonal Carcinogenesis. Springer-Verlag, New York, pp 73–77<br />

275. Chrysogelos SA, Dickson RB 1994 EGF receptor expression, regulation,<br />

and function in breast cancer. Breast Cancer Res Treat<br />

29:29–40<br />

276. Xie W, Paterson AJ, Chin E, Nabell LM, Kudlow JE 1997 Targeted<br />

expression of a dominant negative epidermal growth factor receptor<br />

in the mammary gland of transgenic mice inhibits pubertal<br />

mammary duct development. Mol Endocrinol 11:1766–1781<br />

277. Ankrapp DP, Bennett JM, Haslam SZ 1998 Role of epidermal<br />

growth factor in the acquisition of ovarian steroid hormone responsiveness<br />

in the normal mouse mammary gland. J Cell Physiol<br />

174:251–260<br />

278. Cunha GR, Young P, Hom YK, Cooke PS, Taylor JA, Lubahn DB<br />

1997 Elucidation of a role of stromal steroid hormone receptors in<br />

mammary gland growth and development by tissue recombination<br />

experiments. J Mammary Gland Biol Neoplasia 2:393–402<br />

279. Bole-Feysot C, Goffin V, Edery M, Binart N, Kelly PA 1998 Prolactin<br />

(PRL) and its receptor: actions, signal transduction pathways<br />

and phenotypes observed in PRL receptor knockout mice. Endocr<br />

Rev 19:225–268<br />

280. Chauchereau A, Savouret JF, Milgrom E 1992 Control of biosynthesis<br />

and post-transcriptional modification of the progesterone<br />

receptor. Biol Reprod 46:174–177<br />

281. Maurer RA, Kim KE, Day RN, Notides AC 1990 Regulation of<br />

prolactin gene expression by estradiol. Prog Clin Biol Res 322:159–<br />

169<br />

282. Scully KM, Gleiberman AS, Lindzey J, Lubahn DB, Korach KS,<br />

Rosenfeld MG 1997 Role of estrogen receptor-� in the anterior<br />

pituitary gland. Mol Endocrinol 11:674–681<br />

283. Miller WR, Sharpe RM 1998 Environmental oestrogens and human<br />

reproductive cancers. Endocr Relat Cancer 5:69–96<br />

284. Simpson ER, Zhao Y, Bulun SE, Mendelson CR, Agarwal VR 1997<br />

Mesenchymal-epithelial interactions and breast cancer-role of local<br />

estrogen production. Endocr Relat Cancer 4:407–422<br />

285. Feigelson HS, Henderson BE 1996 <strong>Estrogen</strong>s and breast cancer.<br />

Carcinogenesis 17:2279–2284<br />

286. Collaborative Group on Hormonal Factors in Breast Cancer 1996<br />

Breast cancer and hormonal contraceptives: collaborative reanalysis<br />

of individual data on 53 297 women with breast cancer and 100<br />

239 women without breast cancer from 54 epidemiological studies.<br />

Lancet 347:1713–1727<br />

287. Gotteland M, May E, May-Levin F, Contesso G, Delarue JC,<br />

Mouriesse H 1994 <strong>Estrogen</strong> receptors (ER) in human breast cancer.<br />

The significance of a new prognostic factor based on both ER<br />

protein and ER mRNA contents. Cancer 74:864–871<br />

288. Fanelli MA, Vargas-Roig LM, Gago FE, Tello O, Lucero De Angelis<br />

R, Ciocca DR 1996 <strong>Estrogen</strong> receptors, progesterone receptors,<br />

and cell proliferation in human breast cancer. Breast Cancer<br />

Res Treat 37:217–228<br />

289. Romain S, Chinot O, Guirou O, Soulliere M, Martin PM 1994<br />

Biological heterogeneity of ER-positive breast cancers in the postmenopausal<br />

population. Int J Cancer 59:17–19<br />

290. Mansour EG, Ravdin PM, Dressler L 1994 Prognostic factors in<br />

early breast carcinoma. Cancer 74:381–400<br />

291. Masood S 1992 <strong>Estrogen</strong> and progesterone receptors in cytology:<br />

a comprehensive review. Diagn Cytopathol 8:475–491<br />

292. Simpson ER, Bulun SE, Nichols JE, Zhao Y 1996 <strong>Estrogen</strong> biosynthesis<br />

in adipose tissue: regulation by paracrine and autocrine<br />

mechanisms. J Endocrinol 150[Suppl]:S51–S57<br />

293. Murphy LC, Dotzlaw H, Leygue E, Coutts A, Watson P 1998 The<br />

pathophysiological role of estrogen receptor variants in human<br />

breast cancer. J Steroid Biochem Mol Biol 65:175–180<br />

294. Speirs V, Parkes AT, Kerin MJ, Walton DS, Carleton PJ, Fox JN,<br />

Atkin SL 1999 Coexpression of estrogen receptor-� and -�: poor<br />

prognostic factors in human breast cancer? Cancer Res 59:525–528<br />

295. Nusse R, Varmus HE 1992 Wnt genes. Cell 69:1073–1087<br />

296. Tsukamoto AS, Grosschedl R, Guzman RC, Parslow T, Varmus<br />

HE 1988 Expression of the int-1 gene in transgenic mice is associated<br />

with mammary gland hyperplasia and adenocarcinomas in<br />

male and female mice. Cell 55:619–625<br />

297. Bocchinfuso WP, Hively WP, Couse JF, Varmus HE, Korach KS<br />

1999 An MMTV-Wnt-1 transgene induces mammary gland hyperplasia<br />

and tumorigenesis in mice lacking estrogen receptor-�. Cancer<br />

Res 59:1869–1876<br />

298. Couse JF, Davis VL, Hanson RB, Jefferson WN, McLachlan JA,<br />

Bullock BC, Newbold RR, Korach KS 1997 Accelerated onset of<br />

uterine tumors in transgenic mice with aberrant expression of the<br />

estrogen receptor after neonatal exposure to diethylstilbestrol. Mol<br />

Carcinog 19:236–242<br />

299. Shyamala G, Yang X, Silberstein G, Barcellos-Hoff MH, Dale E 1998<br />

Transgenic mice carrying an imbalance in the native ratio of A to B<br />

forms of progesterone receptor exhibit developmental abnormalities<br />

in mammary glands. Proc Natl Acad Sci USA 95:696–701<br />

300. Gustafson ML, Donahoe PK 1994 Male sex determination: current<br />

concepts of male sexual differentiation. Annu Rev Med 45:505–524<br />

301. Lyon MF, Hawkes SG 1970 X-linked gene for testicular feminization<br />

in the mouse. Nature 227:1217–1219<br />

302. Lindzey J, Kumar MV, Grossman M, Young C, Tindall DJ 1994<br />

Molecular Mechanisms of Androgen Action. In: Litwak G (ed)<br />

Steroids. Academic Press, San Diego, CA, pp 383–432<br />

303. Robaire B, Hermo L 1988 Efferent ducts, epididymis, and vas<br />

deferens: structure, functions, and their regulation. In: Knobil E,


June, 1999 ESTROGEN RECEPTOR NULL MICE 413<br />

Neil JD, Ewing LL, Greenwald GS, Markert CL, Pfaff DW (eds) The<br />

Physiology of Reproduction. Raven Press, New York, pp 999–1080<br />

304. Bartke A, Mason M, Dalterio S, Bex F 1978 Effects of tamoxifen on<br />

plasma concentrations of testosterone and gonadotrophins in the<br />

male rat. J Endocrinol 79:239–240<br />

305. McLachlan JA, Newbold RR, Bullock B 1975 Reproductive tract<br />

lesions in male mice exposed prenatally to diethylstilbestrol. Science<br />

190:991–992<br />

306. Newbold RR, Bullock BC, McLachlan JA 1985 Lesions of the rete<br />

testis in mice exposed prenatally to diethylstilbestrol. Cancer Res<br />

45:5145–5150<br />

307. Newbold RR, Bullock BC, McLachlan JA 1986 Adenocarcinoma<br />

of the rete testis. Diethylstilbestrol-induced lesions of the mouse<br />

rete testis. Am J Pathol 125:625–628<br />

308. Beckman Jr WC, Newbold RR, Teng CT, McLachlan JA 1994<br />

Molecular feminization of mouse seminal vesicle by prenatal exposure<br />

to diethylstilbestrol: altered expression of messenger RNA.<br />

J Urol 151:1370–1378<br />

309. Schleicher G, Drews U, Stumpf WE, Sar M 1984 Differential distribution<br />

of dihydrotestosterone and estradiol binding sites in the<br />

epididymis of the mouse. An autoradiographic study. Histochemistry<br />

81:139–147<br />

310. Cooke PS, Young P, Hess RA, Cunha GR 1991 <strong>Estrogen</strong> receptor<br />

expression in developing epididymis, efferent ductules, and other<br />

male reproductive organs. Endocrinology 128:2874–2879<br />

311. Hess RA, Gist DH, Bunick D, Lubahn DB, Farrell A, Bahr J,<br />

Cooke PS, Greene GL 1997 <strong>Estrogen</strong> receptor (� and �) expression<br />

in the excurrent ducts of the adult male rat reproductive tract. J<br />

Androl 18:602–611<br />

312. van Pelt AM, de Rooij DG, van der Burg B, van der Saag PT,<br />

Gustafsson JÅ, Kuiper GG 1999 Ontogeny of estrogen receptor-�<br />

expression in rat testis. Endocrinology 140:478–483<br />

313. Lau KM, Leav I, Ho SM 1998 Rat estrogen receptor-� and -�, and<br />

progesterone receptor mRNA expression in various prostatic lobes<br />

and microdissected normal and dysplastic epithelial tissues of the<br />

Noble rats. Endocrinology 139:424–427<br />

314. Prins GS, Marmer M, Woodham C, Chang W, Kuiper G, Gustafsson<br />

JÅ, Birch L 1998 <strong>Estrogen</strong> receptor-� messenger ribonucleic<br />

acid ontogeny in the prostate of normal and neonatally estrogenized<br />

rats. Endocrinology 139:874–883<br />

315. Eddy EM, Washburn TF, Bunch DO, Goulding EH, Gladen BC,<br />

Lubahn DB, Korach KS 1996 Targeted disruption of the estrogen<br />

receptor gene in male mice causes alteration of spermatogenesis<br />

and infertility. Endocrinology 137:4796–4805<br />

316. Donaldson KM, Tong SY, Washburn T, Lubahn DB, Eddy EM,<br />

Hutson JM, Korach KS 1996 Morphometric study of the gubernaculum<br />

in male estrogen receptor mutant mice. J Androl 17:91–95<br />

317. Lindzey J, Wetsel WC, Couse JF, Stoker T, Cooper R, Korach KS<br />

1998 Effects of castration and chronic steroid treatments on hypothalamic<br />

gonadotropin-releasing hormone content and pituitary<br />

gonadotropins in male wild-type and estrogen receptor-� knockout<br />

mice. Endocrinology 139:4092–4101<br />

318. Ogawa S, Lubahn DB, Korach KS, Pfaff DW 1997 Behavioral<br />

effects of estrogen receptor gene disruption in male mice. Proc Natl<br />

Acad Sci USA 94:1476–1481<br />

319. Hutson JM, Baker M, Terada M, Zhou B, Paxton G 1994 Hormonal<br />

control of testicular descent and the cause of cryptorchidism. Reprod<br />

Fertil Dev 6:151–156<br />

320. Dix DJ, Allen JW, Collins BW, Mori C, Nakamura N, Poorman-<br />

Allen P, Goulding EH, Eddy EM 1996 Targeted gene disruption<br />

of Hsp70–2 results in failed meiosis, germ cell apoptosis, and male<br />

infertility. Proc Natl Acad Sci USA 93:3264–3268<br />

321. Blendy JA, Kaestner KH, Weinbauer GF, Nieschlag E, Schutz G<br />

1996 Severe impairment of spermatogenesis in mice lacking the<br />

CREM gene. Nature 380:162–165<br />

322. Nantel F, Monaco L, Foulkes NS, Masquilier D, LeMeur M, Henriksen<br />

K, Dierich A, Parvinen M, Sassone-Corsi P 1996 Spermiogenesis<br />

deficiency and germ-cell apoptosis in CREM-mutant mice.<br />

Nature 380:159–162<br />

323. Ruggiu M, Speed R, Taggart M, McKay SJ, Kilanowski F, Saunders<br />

P, Dorin J, Cooke HJ 1997 The mouse Dazla gene encodes a<br />

cytoplasmic protein essential for gametogenesis. Nature 389:73–77<br />

324. Roest HP, van Klaveren J, de Wit J, van Gurp CG, Koken MH,<br />

Vermey M, van Roijen JH, Hoogerbrugge JW, Vreeburg JT,<br />

Baarends WM, Bootsma D, Grootegoed JA, Hoeijmakers JH 1996<br />

Inactivation of the HR6B ubiquitin-conjugating DNA repair enzyme<br />

in mice causes male sterility associated with chromatin modification.<br />

Cell 86:799–810<br />

325. Ritzen EM, Van Damme MP, Froysa B, Reuter C, De La Torre B,<br />

Diczfalusy E 1981 Identification of estradiol produced by Sertoli<br />

cell enriched cultures during incubation with testosterone. J Steroid<br />

Biochem 14:533–535<br />

326. Nakhla AM, Mather JP, Janne OA, Bardin CW 1984 <strong>Estrogen</strong> and<br />

androgen receptors in Sertoli, Leydig, myoid, and epithelial cells:<br />

effects of time in culture and cell density. Endocrinology 115:121–128<br />

327. Hess RA, Bunick D, Lee KH, Bahr J, Taylor JA, Korach KS,<br />

Lubahn DB 1997 A role for oestrogens in the male reproductive<br />

system. Nature 390:509–512<br />

328. Fisher JS, Millar MR, Majdic G, Saunders PT, Fraser HM, Sharpe<br />

RM 1997 Immunolocalisation of oestrogen receptor-� within the<br />

testis and excurrent ducts of the rat and marmoset monkey from<br />

perinatal life to adulthood. J Endocrinol 153:485–495<br />

329. Anton E 1979 Early ultrastructural changes in the rat testis after<br />

ductuli efferentes ligation. Fertil Steril 31:187–194<br />

330. Wang JM, Gu CH, Tao L, Wu XL 1985 Effect of surgery and efferent<br />

duct ligation on testicular blood flow and testicular steroidogenesis<br />

in the rat. J Reprod Fertil 73:191–196<br />

331. Nitta H, Bunick D, Hess RA, Janulis L, Newton SC, Millette CF,<br />

Osawa Y, Shizuta Y, Toda K, Bahr JM 1993 Germ cells of the mouse<br />

testis express P450 aromatase. Endocrinology 132:1396–1401<br />

332. Meistrich ML, Hughes TH, Bruce WR 1975 Alteration of epididymal<br />

sperm transport and maturation in mice by oestrogen and<br />

testosterone. Nature 258:145–147<br />

333. Coffey DS 1988 Androgen action and the sex accessory tissues. In:<br />

Knobil E, Neil JD, Ewing LL, Greenwald GS, Markert CL, Pfaff DW<br />

(eds) The Physiology of Reproduction. Raven Press, New York, pp<br />

1081–1139<br />

334. Prins GS, Woodham C, Lepinske M, Birch L 1993 Effects of neonatal<br />

estrogen exposure on prostatic secretory genes and their<br />

correlation with androgen receptor expression in the separate prostate<br />

lobes of the adult rat. Endocrinology 132:2387–2398<br />

335. Prins GS, Birch L 1995 The developmental pattern of androgen<br />

receptor expression in rat prostate lobes is altered after neonatal<br />

exposure to estrogen. Endocrinology 136:1303–1314<br />

336. Prins GS, Birch L 1997 Neonatal estrogen exposure up-regulates<br />

estrogen receptor expression in the developing and adult rat prostate<br />

lobes. Endocrinology 138:1801–189<br />

337. McEwen BS 1992 Steroid hormones: effect on brain development<br />

and function. Horm Res 37:1–10<br />

338. Blaustein JD 1994 <strong>Estrogen</strong> receptors in neurons: new subcellular<br />

locations and functional implications. <strong>Endocrine</strong> 2:249–258<br />

339. Stefaneanu L 1997 Pituitary sex steroid receptors: localization and<br />

function. Endocr Pathol 8:91–108<br />

340. Smith SS 1994 Female sex steroid hormones: from receptors to<br />

networks to performance–actions on the sensorimotor system. Prog<br />

Neurobiol 44:55–86<br />

341. Backstrom T 1992 Neuroendocrinology of premenstrual syndrome.<br />

Clin Obstet Gynecol 35:612–628<br />

342. McEwen BS, Alves SE, Bulloch K, Weiland NG 1998 Clinically<br />

relevant basic science studies of gender differences and sex hormone<br />

effects. Psychopharmacol Bull 34:251–259<br />

343. Kawata M 1995 Roles of steroid hormones and their receptors in<br />

structural organization in the nervous system. Neurosci Res 24:1–46<br />

344. Spindler KD 1997 Interactions between steroid hormones and the<br />

nervous system. Neurotoxicology 18:745–754<br />

345. Sachs BD, Meisel RL 1988 The physiology of male sexual behavior.<br />

In: Knobil E, Neil JD, Ewing LL, Greenwald GS, Markert CL, Pfaff<br />

DW (eds) The Physiology of Reproduction. Raven Press, New York,<br />

pp 1393–1485<br />

346. MacLusky NJ, Leiberburg I, McEwen BS 1979 Development of<br />

steroid receptor systems in rodent brain. In: Hamilton TH, Clark<br />

JH, Sadler WA (eds) Ontogeny of <strong>Receptor</strong>s and Reproductive<br />

Hormone Action. Raven Press, New York, pp 393–402<br />

347. McEwen BS 1992 Effects of the steroid/thyroid hormone family on<br />

neural and behavioral plasticity. In: Nemeroff CB (ed) Neuroendocrinology.<br />

CRC Press, Boca Raton, FL, pp 333–351


414 COUSE AND KORACH Vol. 20, No. 3<br />

348. Gorski RA 1979 Nature of hormone action in the brain. In: Hamilton<br />

TH, Clark JH, Sadler WA (eds) Ontogeny of <strong>Receptor</strong>s and Reproductive<br />

Hormone Action. Raven Press, New York, pp 371–392<br />

349. Pilgrim C, Hutchison JB 1994 Developmental regulation of sex<br />

differences in the brain: can the role of gonadal steroids be redefined?<br />

Neuroscience 60:843–855<br />

350. Shughrue PJ, Lane MV, Merchenthaler I 1997 Comparative distribution<br />

of estrogen receptor-� and -� mRNA in the rat central<br />

nervous system. J Comp Neurol 388:507–525<br />

351. Laflamme N, Nappi RE, Drolet G, Labrie C, Rivest S 1998 Expression<br />

and neuropeptidergic characterization of estrogen receptors<br />

(ER� and ER�) throughout the rat brain: anatomical evidence<br />

of distinct roles of each subtype. J Neurobiol 36:357–378<br />

352. Shughrue P, Scrimo P, Lane M, Askew R, Merchenthaler I 1997<br />

The distribution of estrogen receptor-� mRNA in forebrain regions<br />

of the estrogen receptor-� knockout mouse. Endocrinology 138:<br />

5649–5652<br />

353. Osterlund M, Kuiper GG, Gustafsson JÅ, Hurd YL 1998 Differential<br />

distribution and regulation of estrogen receptor-� and -�<br />

mRNA within the female rat brain. Brain Res Mol Brain Res 54:<br />

175–180<br />

354. Register TC, Shively CA, Lewis CE 1998 Expression of estrogen<br />

receptor-� and -� transcripts in female monkey hippocampus and<br />

hypothalamus. Brain Res 788:320–322<br />

355. Wehling M 1997 Specific, nongenomic actions of steroid hormones.<br />

Annu Rev Physiol 59:365–393<br />

356. McEwen BS 1991 Non-genomic and genomic effects of steroids on<br />

neural activity. Trends Pharmacol Sci 12:141–147<br />

357. Wong M, Thompson TL, Moss RL 1996 Nongenomic actions of<br />

estrogen in the brain: physiological significance and cellular mechanisms.<br />

Crit Rev Neurobiol 10:189–203<br />

358. Wiebe JP 1997 Nongenomic actions of steroids on gonadotropin<br />

release. Recent Prog Horm Res 52:71–101<br />

359. Pietras RJ, Szego CM 1977 Specific binding sites for oestrogen at<br />

the outer surfaces of isolated endometrial cells. Nature 265:69–72<br />

360. Watson CS, Pappas TC, Gametchu B 1995 The other estrogen receptor<br />

in the plasma membrane: implications for the actions of environmental<br />

estrogens. Environ Health Perspect 103[Suppl 7]:41–50<br />

361. Karthikeyan N, Thampan RV 1996 Plasma membrane is the primary<br />

site of localization of the nonactivated estrogen receptor in the<br />

goat uterus: hormone binding causes receptor internalization. Arch<br />

Biochem Biophys 325:47–57<br />

362. Pappas TC, Gametchu B, Watson CS 1995 Membrane estrogen<br />

receptors identified by multiple antibody labeling and impededligand<br />

binding. FASEB J 9:404–410<br />

363. Razandi M, Pedram A, Greene GL, Levin ER 1999 Cell membrane<br />

and nuclear estrogen receptors (ERs) originate from a single transcript:<br />

studies of ER� and ER� expressed in Chinese hamster ovary<br />

cells. Mol Endocrinol 13:307–319<br />

364. Gu Q, Korach KS, Moss RL 1999 Rapid action of 17�-estradiol on<br />

kainate-induced currents in hippocampal neurons lacking intracellular<br />

estrogen receptors. Endocrinology 140:660–666<br />

365. Karsch FJ 1984 The hypothalamus and anterior pituitary gland. In:<br />

Austin CR, Short RV (eds) Reproduction in Mammals: Hormonal<br />

Control of Reproduction. Cambridge University Press, Cambridge,<br />

UK, pp 1–20<br />

366. Stefaneanu L, Kovacs K, Horvath E, Lloyd RV, Buchfelder M,<br />

Fahlbusch R, Smyth H 1994 In situ hybridization study of estrogen<br />

receptor messenger ribonucleic acid in human adenohypophysial<br />

cells and pituitary adenomas. J Clin Endocrinol Metab 78:83–88<br />

367. Sar M, Parikh I 1986 Immunohistochemical localization of estrogen<br />

receptor in rat brain, pituitary and uterus with monoclonal antibodies.<br />

J Steroid Biochem 24:497–503<br />

368. Friend KE, Resnick EM, Ang LW, Shupnik MA 1997 Specific<br />

modulation of estrogen receptor mRNA isoforms in rat pituitary<br />

throughout the estrous cycle and in response to steroid hormones.<br />

Mol Cell Endocrinol 131:147–155<br />

369. Shupnik MA, Gordon MS, Chin WW 1989 Tissue-specific regulation<br />

of rat estrogen receptor mRNAs. Mol Endocrinol 3:660–665<br />

370. Korach KS, Muldoon TG 1973 Comparison of specific 17�-estradiol-receptor<br />

interactions in the anterior pituitary of male and<br />

female rats. Endocrinology 92:322–326<br />

371. Notides AC 1970 The binding affinity and specificity of the estro-<br />

gen receptor of the rat uterus and anterior pituitary. Endocrinology<br />

87:987–992<br />

372. Shupnik MA 1996 Gonadotropin gene modulation by steroids and<br />

gonadotropin-releasing hormone. Biol Reprod 54:279–286<br />

373. Gharib SD, Wierman ME, Shupnik MA, Chin WW 1990 Molecular<br />

biology of the pituitary gonadotropins. Endocr Rev 11:177–199<br />

374. Belchetz PE, Plant TM, Nakai Y, Keogh EJ, Knobil E 1978 Hypophysial<br />

responses to continuous and intermittent delivery of hypopthalamic<br />

gonadotropin-releasing hormone. Science 202:631–633<br />

375. Veldhuis JD 1990 The hypothalamic pulse generator: the reproductive<br />

core. Clin Obstet Gynecol 33:538–550<br />

376. Haisenleder DJ, Dalkin AC, Marshall JC 1994 Regulation of gonadotropin<br />

gene expression. In: Knobil E, Neil JD, Ewing LL,<br />

Greenwald GS, Markert CL, Pfaff DW (eds) The Physiology of<br />

Reproduction. Raven Press, New York, pp 1793–1813<br />

377. Fink G 1988 Gonadotropin secretion and its control. In: Knobil E,<br />

Neil JD, Ewing LL, Greenwald GS, Markert CL, Pfaff DW (eds) The<br />

Physiology of Reproduction. Raven Press, New York, pp 1349–1377<br />

378. Wise PM, Ratner A 1980 Effect of ovariectomy on plasma LH, FSH,<br />

estradiol, and progesterone and medial basal hypothalamic LHRH<br />

concentrations old and young rats. Neuroendocrinology 30:15–19<br />

379. Yamamoto M, Diebel ND, Bogdanove EM 1970 Analysis of initial<br />

and delayed effects of orchidectomy and ovariectomy on pituitary<br />

and serum LH levels in adult and immature rats. Endocrinology<br />

86:1102–1111<br />

380. Sarkar DK, Fink G 1980 Luteinizing hormone releasing factor in<br />

pituitary stalk plasma from long-term ovariectomized rats: effects<br />

of steroids. J Endocrinol 86:511–524<br />

381. Levine JE, Ramirez VD 1982 Luteinizing hormone-releasing hormone<br />

release during the rat estrous cycle and after ovariectomy, as<br />

estimated with push-pull cannulae. Endocrinology 111:1439–1448<br />

382. Rodin DA, Lalloz MR, Clayton RN 1989 Gonadotropin-releasing<br />

hormone regulates follicle-stimulating hormone �-subunit gene<br />

expression in the male rat. Endocrinology 125:1282–1289<br />

383. Lalloz MR, Detta A, Clayton RN 1988 Gonadotropin-releasing<br />

hormone is required for enhanced luteinizing hormone subunit<br />

gene expression in vivo. Endocrinology 122:1681–1688<br />

384. Keri RA, Andersen B, Kennedy GC, Hamernik DL, Clay CM,<br />

Brace AD, Nett TM, Notides AC, Nilson JH 1991 Estradiol inhibits<br />

transcription of the human glycoprotein hormone �-subunit gene<br />

despite the absence of a high-affinity binding site for estrogen<br />

receptor. Mol Endocrinol 5:725–733<br />

385. Fallest PC, Trader GL, Darrow JM, Shupnik MA 1995 Regulation<br />

of rat luteinizing hormone � gene expression in transgenic mice by<br />

steroids and a gonadotropin-releasing hormone antagonist. Biol<br />

Reprod 53:103–109<br />

386. Keri RA, Wolfe MW, Saunders TL, Anderson I, Kendall SK,<br />

Wagner T, Yeung J, Gorski J, Nett TM, Camper SA, Nilson JH<br />

1994 The proximal promoter of the bovine luteinizing hormone<br />

�-subunit gene confers gonadotrope-specific expression and regulation<br />

by gonadotropin-releasing hormone, testosterone, and 17<br />

�-estradiol in transgenic mice. Mol Endocrinol 8:1807–1816<br />

387. Chappell PE, Lydon JP, Conneely OM, O’Malley BW, Levine JE<br />

1997 <strong>Endocrine</strong> defects in mice carrying a null mutation for the<br />

progesterone receptor gene. Endocrinology 138:4147–4152<br />

388. Vale W, Bilezikjian LM, Rivier C 1994 Reproductive and other<br />

roles of inhibins and activins. In: Knobil E, Neil JD, Ewing LL,<br />

Greenwald GS, Markert CL, Pfaff DW (eds) The Physiology of<br />

Reproduction. Raven Press, New York, pp 1861–1878<br />

389. Wang QF, Farnworth PG, Findlay JK, Burger HG 1988 Effect of<br />

purified 31K bovine inhibin on the specific binding of gonadotropin-releasing<br />

hormone to rat anterior pituitary cells in culture.<br />

Endocrinology 123:2161–2166<br />

390. Carroll RS, Corrigan AZ, Gharib SD, Vale W, Chin WW 1989<br />

Inhibin, activin, and follistatin: regulation of follicle-stimulating<br />

hormone messenger ribonucleic acid levels. Mol Endocrinol<br />

3:1969–1976<br />

391. Attardi B, Keeping HS, Winters SJ, Kotsuji F, Troen P 1991 Comparison<br />

of the effects of cycloheximide and inhibin on the gonadotropin<br />

subunit messenger ribonucleic acids. Endocrinology 128:<br />

119–125<br />

392. Mercer JE, Clements JA, Funder JW, Clarke IJ 1987 Rapid and


June, 1999 ESTROGEN RECEPTOR NULL MICE 415<br />

specific lowering of pituitary FSH� mRNA levels by inhibin. Mol<br />

Cell Endocrinol 53:251–254<br />

393. Fukuda M, Miyamoto K, Hasegawa Y, Ibuki Y, Igarashi M 1987<br />

Action mechanism of inhibin in vitro–cycloheximide mimics inhibin<br />

actions on pituitary cells. Mol Cell Endocrinol 51:41–50<br />

394. Attardi B, Miklos J 1990 Rapid stimulatory effect of activin-A on<br />

messenger RNA encoding the follicle-stimulating hormone �-subunit<br />

in rat pituitary cell cultures. Mol Endocrinol 4:721–726<br />

395. Carroll RS, Corrigan AZ, Vale W, Chin WW 1991 Activin stabilizes<br />

follicle-stimulating hormone-� messenger ribonucleic acid levels.<br />

Endocrinology 129:1721–1726<br />

396. Kaiser UB, Jakubowiak A, Steinberger A, Chin WW 1993 Regulation<br />

of rat pituitary gonadotropin-releasing hormone receptor<br />

mRNA levels in vivo and in vitro. Endocrinology 133:931–934<br />

397. Kaiser UB, Sabbagh E, Katzenellenbogen RA, Conn PM, Chin<br />

WW 1995 A mechanism for the differential regulation of gonadotropin<br />

subunit gene expression by gonadotropin-releasing hormone.<br />

Proc Natl Acad Sci USA 92:12280–12284<br />

398. Nakai Y, Plant TM, Hess DL, Keogh EJ, Knobil E 1978 On the sites<br />

of the negative and positive feedback actions of estradiol in the<br />

control of gonadotropin secretion in the rhesus monkey. Endocrinology<br />

102:1008–1014<br />

399. Shupnik MA, Weinmann CM, Notides AC, Chin WW 1989 An<br />

upstream region of the rat luteinizing hormone � gene binds estrogen<br />

receptor and confers estrogen responsiveness. J Biol Chem<br />

264:80–86<br />

400. Shughrue PJ, Lubahn DB, Negro-Vilar A, Korach KS, Merchenthaler<br />

I 1997 Responses in the brain of estrogen receptor �-disrupted<br />

mice. Proc Natl Acad Sci USA 94:11008–11012<br />

401. Clark CR, MacLusky NJ, Naftolin F 1982 Oestrogen induction of<br />

progestin receptors in the rat brain and pituitary gland: quantitative<br />

and kinetic aspects. J Endocrinol 93:339–353<br />

402. Simerly RB, Swanson LW, Gorski RA 1985 Reversal of the sexually<br />

dimorphic distribution of serotonin-immunoreactive fibers in<br />

the medial preoptic nucleus by treatment with perinatal androgen.<br />

Brain Res 340:91–98<br />

403. Simerly RB 1989 Hormonal control of the development and regulation<br />

of tyrosine hydroxylase expression within a sexually dimorphic<br />

population of dopaminergic cells in the hypothalamus.<br />

Brain Res Mol Brain Res 6:297–310<br />

404. Simerly RB, Swanson LW, Handa RJ, Gorski RA 1985 Influence<br />

of perinatal androgen on the sexually dimorphic distribution of<br />

tyrosine hydroxylase-immunoreactive cells and fibers in the anteroventral<br />

periventricular nucleus of the rat. Neuroendocrinology<br />

40:501–510<br />

405. Simerly RB, Zee MC, Pendleton JW, Lubahn DB, Korach KS 1997<br />

<strong>Estrogen</strong> receptor-dependent sexual differentiation of dopaminergic<br />

neurons in the preoptic region of the mouse. Proc Natl Acad Sci<br />

USA 94:14077–14082<br />

406. Lephart ED 1996 A review of brain aromatase cytochrome P450.<br />

Brain Res Rev 22:1–26<br />

407. Negri-Cesi P, Poletti A, Celotti F 1996 Metabolism of steroids in<br />

the brain: a new insight into the role of 5�-reductase and aromatase<br />

in brain differentiation and functions. J Steroid Biochem Mol Biol<br />

58:455–466<br />

408. Day RN, Koike S, Sakai M, Muramatsu M, Maurer RA 1990 Both<br />

Pit-1 and the estrogen receptor are required for estrogen responsiveness<br />

of the rat prolactin gene. Mol Endocrinol 4:1964–1971<br />

409. Nowakowski BE, Maurer RA 1994 Multiple Pit-1-binding sites<br />

facilitate estrogen responsiveness of the prolactin gene. Mol Endocrinol<br />

8:1742–1749<br />

410. Frawley LS, Boockfor FR 1991 Mammosomatotropes: presence<br />

and functions in normal and neoplastic pituitary tissue. Endocr Rev<br />

12:337–355<br />

411. Slabaugh MB, Lieberman ME, Rutledge JJ, Gorski J 1982 Ontogeny<br />

of growth hormone and prolactin gene expression in mice.<br />

Endocrinology 110:1489–1497<br />

412. Gash D, Ahmad N, Weiner R, Schechter J 1980 Development of<br />

presumptive mammotrophs in isografts is dependent on the endocrine<br />

state of the host. Endocrinology 106:1246–1252<br />

413. Chun TY, Gregg D, Sarkar DK, Gorski J 1998 Differential regulation<br />

by estrogens of growth and prolactin synthesis in pituitary<br />

cells suggests that only a small pool of estrogen receptors is required<br />

for growth. Proc Natl Acad Sci USA 95:2325–2330<br />

414. Baum MJ 1979 Differentiation of coital behavior in mammals: a<br />

comparative analysis. Neurosci Biobehav Rev 3:265–284<br />

415. McCarthy MM, Schlenker EH, Pfaff DW 1993 Enduring consequences<br />

of neonatal treatment with antisense oligodeoxynucleotides<br />

to estrogen receptor messenger ribonucleic acid on sexual<br />

differentiation of rat brain. Endocrinology 133:433–439<br />

416. Pfaff DW, Scwartz-Giblin S, McCarthy MM, Kow LM 1994 Cellular<br />

and molecular mechanisms of female reproductive behaviors.<br />

In: Knobil E, Neil JD, Ewing LL, Greenwald GS, Markert CL, Pfaff<br />

DW (eds) The Physiology of Reproduction. Raven Press, New York,<br />

pp 107–220<br />

417. Ogawa S, Taylor JA, Lubahn DB, Korach KS, Pfaff DW 1996<br />

Reversal of sex roles in genetic female mice by disruption of estrogen<br />

receptor gene. Neuroendocrinology 64:467–470<br />

418. Ogawa S, Eng V, Taylor J, Lubahn DB, Korach KS, Pfaff DW 1998<br />

Roles of estrogen receptor-� gene expression in reproduction-related<br />

behaviors in female mice. Endocrinology 139:5070–5081<br />

419. Rissman EF, Early AH, Taylor JA, Korach KS, Lubahn DB 1997<br />

<strong>Estrogen</strong> receptors are essential for female sexual receptivity. Endocrinology<br />

138:507–510<br />

420. Boling JL, Blandau RH 1939 The estrogen-progesterone induction<br />

of mating responses in the spayed female rat. Endocrinology 25:<br />

359–364<br />

421. Blaustein JD, Feder HH 1979 Cytoplasmic progestin receptors in<br />

female guinea pig brain and their relationship to refractoriness in<br />

expression of female sexual behavior. Brain Res 177:489–498<br />

422. Blaustein JD, Feder HH 1980 Nuclear progestin receptors in guinea<br />

pig brain measured by an in vitro exchange assay after hormonal<br />

treatments that affect lordosis. Endocrinology 106:1061–1069<br />

423. Moffatt CA, Rissman EF, Shupnik MA, Blaustein JD 1998 Induction<br />

of progestin receptors by estradiol in the forebrain of estrogen<br />

receptor-� gene-disrupted mice. J Neurosci 18:9556–9563<br />

424. Simon NG, Whalen RE, Tate MP 1985 Induction of male-typical<br />

aggression by androgens but not by estrogens in adult female mice.<br />

Horm Behav 19:204–212<br />

425. Simon NG, Whalen RE 1987 Sexual differentiation of androgensensitive<br />

and estrogen-sensitive regulatory systems for aggressive<br />

behavior. Horm Behav 21:493–500<br />

426. Ogawa S, Krebs CJ, Korach KS, Pfaff DW, Forebrain steroid<br />

receptor immunoreactive cells in neonatal and prepubertal estrogen<br />

receptor-� gene deficient (ERKO) mice. Program of the 28th<br />

Annual Meeting of the Society for Neuroscience, Los Angeles, CA,<br />

1998 (Abstract 81.7), p 199<br />

427. Wersinger SR, Sannen K, Villalba C, Lubahn DB, Rissman EF, De<br />

Vries GJ 1997 Masculine sexual behavior is disrupted in male and<br />

female mice lacking a functional estrogen receptor � gene. Horm<br />

Behav 32:176–183<br />

428. Ogawa S, Washburn TF, Taylor J, Lubahn DB, Korach KS, Pfaff<br />

DW 1998 Modifications of testosterone-dependent behaviors by<br />

estrogen receptor-� gene disruption in male mice. Endocrinology<br />

139:5058–5069<br />

429. Ono S, Geller LN, Lai EV 1974 TfM mutation and masculinization vs.<br />

feminization of the mouse central nervous system. Cell 3:235–242<br />

430. Ogawa S, Lubahn DB, Korach KS, Pfaff DW 1996 Aggressive<br />

behaviors of transgenic estrogen-receptor knockout male mice.<br />

Ann NY Acad Sci 794:384–385<br />

431. Ciocca DR, Roig LM 1995 <strong>Estrogen</strong> receptors in human nontarget<br />

tissues: biological and clinical implications. Endocr Rev 16:35–62<br />

432. Albright F, Smith PH, Richardson AM 1941 Postmenopausal osteoporosis.<br />

JAMA 116:2465–2474<br />

433. Barrett-Connor E, Grady D 1998 Hormone replacement therapy,<br />

heart disease, and other considerations. Annu Rev Public Health<br />

19:55–72<br />

434. Komm BS, Bodine PV 1998 The ongoing saga of osteoporosis<br />

treatment. J Cell Biochem Suppl 31:277–283<br />

435. Prestwood KM, Pilbeam CC, Raisz LG 1995 Treatment of osteoporosis.<br />

Annu Rev Med 46:249–256<br />

436. McDonnell DP, Norris JD 1997 Analysis of the molecular pharmacology<br />

of estrogen receptor agonists and antagonists provides<br />

insights into the mechanism of action of estrogen in bone. Osteoporos<br />

Int 7:S29–S34


416 COUSE AND KORACH Vol. 20, No. 3<br />

437. Komm BS, Terpening CM, Benz DJ, Graeme KA, Gallegos A,<br />

Korc M, Greene GL, O’Malley BW, Haussler MR 1988 <strong>Estrogen</strong><br />

binding, receptor mRNA, and biologic response in osteoblast-like<br />

osteosarcoma cells. Science 241:81–84<br />

438. Eriksen EF, Colvard DS, Berg NJ, Graham ML, Mann KG, Spelsberg<br />

TC, Riggs BL 1988 Evidence of estrogen receptors in normal<br />

human osteoblast-like cells. Science 241:84–86<br />

439. Davis VL, Couse JF, Gray TK, Korach KS 1994 Correlation between<br />

low levels of estrogen receptors and estrogen responsiveness<br />

in two rat osteoblast-like cell lines. J Bone Miner Res 9:983–991<br />

440. Oursler MJ, Osdoby P, Pyfferoen J, Riggs BL, Spelsberg TC 1991<br />

Avian osteoclasts as estrogen target cells. Proc Natl Acad Sci USA<br />

88:6613–6617<br />

441. Hoyland JA, Mee AP, Baird P, Braidman IP, Mawer EB, Freemont<br />

AJ 1997 Demonstration of estrogen receptor mRNA in bone using in<br />

situ reverse-transcriptase polymerase chain reaction. Bone 20:87–92<br />

442. Stulc T, Klement D, Kvasnicka J, Stepan JJ 1997 Immunocytochemical<br />

detection of estrogen receptors in bone cells using flow<br />

cytometry. Biochim Biophys Acta 1356:95–100<br />

443. Bodine PV, Henderson RA, Green J, Aronow M, Owen T, Stein<br />

GS, Lian JB, Komm BS 1998 <strong>Estrogen</strong> receptor-� is developmentally<br />

regulated during osteoblast differentiation and contributes to<br />

selective responsiveness of gene expression. Endocrinology 139:<br />

2048–2057<br />

444. Onoe Y, Miyaura C, Ohta H, Nozawa S, Suda T 1997 Expression<br />

of estrogen receptor-� in rat bone. Endocrinology 138:4509–4512<br />

445. Turner RT, Riggs BL, Spelsberg TC 1994 Skeletal effects of estrogen.<br />

Endocr Rev 15:275–300<br />

446. Vanderschueren D, Van Herck E, Suiker AM, Visser WJ, Schot<br />

LP, Chung K, Lucas RS, Einhorn TA, Bouillon R 1993 Bone and<br />

mineral metabolism in the androgen-resistant (testicular feminized)<br />

male rat. J Bone Miner Res 8:801–809<br />

447. Korach KS, Taki M, Kimbro KS 1997 The effects of estrogen<br />

receptor gene disruption on bone. In: Paoletti R (ed) Women’s<br />

Health and Menopause. Kluwer Academic Publishers and Fondazione<br />

Giovanni Lorenzini, Amsterdam, The Netherlands, pp 69–73<br />

448. Pan LC, Ke HZ, Simmons HA, Crawford DT, ChidseyFrink KL,<br />

McCurdy SP, Schafer JR, Kimbro KS, Taki M, Korach KS,<br />

Thompson DD 1997 <strong>Estrogen</strong> receptor-� knockout (ERKO) mice<br />

lose trabecular and cortical bone following ovariectomy. J Bone<br />

Miner Res 12:126–126<br />

449. Migliaccio S, Newbold RR, Bullock BC, McLachlan JA, Korach<br />

KS 1992 Developmental exposure to estrogens induces persistent<br />

changes in skeletal tissue. Endocrinology 130:1756–1758<br />

450. Migliaccio S, Newbold RR, Bullock BC, Jefferson WJ, Sutton Jr<br />

FG, McLachlan JA, Korach KS 1996 Alterations of maternal estrogen<br />

levels during gestation affect the skeleton of female offspring.<br />

Endocrinology 137:2118–2125<br />

451. Kusec V, Virdi AS, Prince R, Triffitt JT 1998 Localization of<br />

estrogen receptor-� in human and rabbit skeletal tissues. J Clin<br />

Endocrinol Metab 83:2421–2428<br />

452. Nilsson LO, Boman A, Savendahl L, Grigelioniene G, Ohlsson C,<br />

Ritzen EM, Wroblewski J 1999 Demonstration of estrogen receptor-�<br />

immunoreactivity in human growth plate cartilage. J Clin<br />

Endocrinol Metab 84:370–373<br />

453. Barrett-Connor E 1994 Heart disease in women. Fertil Steril 62:<br />

127S–132S<br />

454. Nathan L, Chaudhuri G 1997 <strong>Estrogen</strong>s and atherosclerosis. Annu<br />

Rev Pharmacol Toxicol 37:477–515<br />

455. Lundeen SG, Carver JM, McKean ML, Winneker RC 1997 Characterization<br />

of the ovariectomized rat model for the evaluation of<br />

estrogen effects on plasma cholesterol levels. Endocrinology 138:<br />

1552–1558<br />

456. Shimano H, Yamada N, Katsuki M, Shimada M, Gotoda T,<br />

Harada K, Murase T, Fukazawa C, Takaku F, Yazaki Y 1992<br />

Overexpression of apolipoprotein E in transgenic mice: marked<br />

reduction in plasma lipoproteins except high density lipoprotein<br />

and resistance against diet-induced hypercholesterolemia. Proc<br />

Natl Acad Sci USA 89:1750–1754<br />

457. Piedrahita JA, Zhang SH, Hagaman JR, Oliver PM, Maeda N 1992<br />

Generation of mice carrying a mutant apolipoprotein E gene inactivated<br />

by gene targeting in embryonic stem cells. Proc Natl Acad<br />

Sci USA 89:4471–4475<br />

458. Staels B, Auwerx J, Chan L, van Tol A, Rosseneu M, Verhoeven<br />

G 1989 Influence of development, estrogens, and food intake on<br />

apolipoprotein A-I, A-II, and E mRNA in rat liver and intestine. J<br />

Lipid Res 30:1137–1145<br />

459. Seishima M, Bisgaier CL, Davies SL, Glickman RM 1991 Regulation<br />

of hepatic apolipoprotein synthesis in the 17�-ethinyl estradiol-treated<br />

rat. J Lipid Res 32:941–951<br />

460. Srivastava RA, Srivastava N, Averna M, Lin RC, Korach KS,<br />

Lubahn DB, Schonfeld G 1997 <strong>Estrogen</strong> up-regulates apolipoprotein<br />

E (ApoE) gene expression by increasing ApoE mRNA in the<br />

translating pool via the estrogen receptor �-mediated pathway.<br />

J Biol Chem 272:33360–33366<br />

461. White MM, Zamudio S, Stevens T, Tyler R, Lindenfeld J, Leslie<br />

K, Moore LG 1995 <strong>Estrogen</strong>, progesterone, and vascular reactivity:<br />

potential cellular mechanisms. Endocr Rev 16:739–751<br />

462. Mendelsohn ME, Karas RH 1994 <strong>Estrogen</strong> and the blood vessel<br />

wall. Curr Opin Cardiol 9:619–626<br />

463. Losordo DW, Kearney M, Kim EA, Jekanowski J, Isner JM 1994<br />

Variable expression of the estrogen receptor in normal and atherosclerotic<br />

coronary arteries of premenopausal women. Circulation<br />

89:1501–1510<br />

464. Iafrati MD, Karas RH, Aronovitz M, Kim S, Sullivan Jr TR,<br />

Lubahn DB, O’Donnell Jr TF, Korach KS, Mendelsohn ME 1997<br />

<strong>Estrogen</strong> inhibits the vascular injury response in estrogen receptor<br />

�-deficient mice. Nat Med 3:545–548<br />

465. Register TC, Adams MR 1998 Coronary artery and cultured aortic<br />

smooth muscle cells express mRNA for both the classical estrogen<br />

receptor and the newly described estrogen receptor-�. J Steroid<br />

Biochem Mol Biol 64:187–191<br />

466. Tschugguel W, Schneeberger C, Zhegu Z, Wieser F, Waselmayr<br />

B, Sator MO, Wojita J, Binder BR, Huber JC 1998 Distinct expression<br />

pattern of estrogen receptor-� and -� mRNA in cultured<br />

human endothelial and vascular smooth muscle cells. J Soc Gynecol<br />

Invest 5 [Suppl]:143A (Abstract)<br />

467. Rubanyi GM, Freay AD, Kauser K, Sukovich D, Burton G,<br />

Lubahn DB, Couse JF, Curtis SW, Korach KS 1997 Vascular estrogen<br />

receptors and endothelium-derived nitric oxide production<br />

in the mouse aorta. Gender difference and effect of estrogen receptor<br />

gene disruption. J Clin Invest 99:2429–2437<br />

468. Paigen B, Holmes PA, Mitchell D, Albee D 1987 Comparison of<br />

atherosclerotic lesions and HDL-lipid levels in male, female, and<br />

testosterone-treated female mice from strains C57BL/6, BALB/c,<br />

and C3H. Atherosclerosis 64:215–221<br />

469. Sullivan Jr TR, Karas RH, Aronovitz M, Faller GT, Ziar JP, Smith<br />

JJ, O’Donnell Jr TF, Mendelsohn ME 1995 <strong>Estrogen</strong> inhibits the<br />

response-to-injury in a mouse carotid artery model. J Clin Invest<br />

96:2482–2488<br />

470. Grohe C, Kahlert S, Lobbert K, Meyer R, Linz KW, Karas RH,<br />

Vetter H 1996 Modulation of hypertensive heart disease by estrogen.<br />

Steroids 61:201–204<br />

471. Grohe C, Kahlert S, Lobbert K, Vetter H 1998 Expression of oestrogen<br />

receptor-� and -� in rat heart: role of local oestrogen synthesis.<br />

J Endocrinol 156:R1–R7<br />

472. Johnson BD, Zheng W, Korach KS, Scheuer T, Catterall WA,<br />

Rubanyi GM 1997 Increased expression of the cardiac L-type calcium<br />

channel in estrogen receptor-deficient mice. J Gen Physiol<br />

110:135–140<br />

473. Gimble JM, Robinson CE, Clarke SL, Hill MR 1998 Nuclear<br />

hormone receptors and adipogenesis. Crit Rev Eukaryot Gene Expr<br />

8:141–168<br />

474. Price TM, O’Brien SN, Welter BH, George R, Anandjiwala J,<br />

Kilgore M 1998 <strong>Estrogen</strong> regulation of adipose tissue lipoprotein<br />

lipase–possible mechanism of body fat distribution. Am J Obstet<br />

Gynecol 178:101–107<br />

475. Frisch RE 1987 Body fat, menarche, fitness and fertility. Hum<br />

Reprod 2:521–533<br />

476. Sibonga JD, Dobnig H, Harden RM, Turner RT 1998 Effect of the<br />

high-affinity estrogen receptor ligand ICI 182,780 on the rat tibia.<br />

Endocrinology 139:3736–3742<br />

477. Turner CH, Sato M, Bryant HU 1994 Raloxifene preserves bone<br />

strength and bone mass in ovariectomized rats. Endocrinology<br />

135:2001–2005<br />

478. Lyon MF, Glenister PH 1980 Reduced reproductive performance


June, 1999 ESTROGEN RECEPTOR NULL MICE 417<br />

in androgen-resistant Tfm/Tfm female mice. Proc R Soc Lond B<br />

Biol Sci 208:1–12<br />

479. Pedersen SB, Borglum JD, Moller-Pedersen T, Richelsen B 1992<br />

Effects of in vivo estrogen treatment on adipose tissue metabolism<br />

and nuclear estrogen receptor binding in isolated rat adipocytes.<br />

Mol Cell Endocrinol 85:13–19<br />

480. Price TM, O’Brien SN 1993 Determination of estrogen receptor<br />

messenger ribonucleic acid (mRNA) and cytochrome P450 aromatase<br />

mRNA levels in adipocytes and adipose stromal cells by<br />

competitive polymerase chain reaction amplification. J Clin Endocrinol<br />

Metab 77:1041–1045<br />

481. Day JK, Bagegni A, MacDonald RS, Lubahn DB, The phytoestrogen,<br />

genestein, inhibits growth in an ER�KO mouse model via an<br />

estrogen receptor-� dependent mechanism. Program of the 80th<br />

Annual Meeting of The <strong>Endocrine</strong> Society, New Orleans, LA, 1998<br />

(Abstract P1-20), p 126<br />

482. Setchell KD 1998 Phytoestrogens: the biochemistry, physiology,<br />

and implications for human health of soy isoflavones. Am J Clin<br />

Nutr 68:1333S–1346S<br />

483. Scambia G, Ferrandina G, D’Agostino G, Fagotti A, Di Stefano M,<br />

Fanfani F, Serri FG, Mancuso S 1998 Oestrogen and progesterone<br />

receptors in ovarian carcinoma. Endocr Relat Cancer 5:293–301<br />

484. Clinton GM, Hua W 1997 <strong>Estrogen</strong> action in human ovarian cancer.<br />

Crit Rev Oncol Hematol 25:1–9<br />

485. Rao BR, Slotman BJ 1991 <strong>Endocrine</strong> factors in common epithelial<br />

ovarian cancer. Endocr Rev 12:14–26<br />

486. Yen S 1990 Clinical endocrinology of reproduction. In: Baulieu EE,<br />

Kelly PA (eds) Hormones: From Molecules to Disease. Hermann,<br />

Chapman and Hall, Paris, pp 445–481<br />

487. Franks S 1989 Polycystic ovary syndrome: a changing perspective.<br />

Clin Endocrinol (Oxf) 31:87–120<br />

488. Kahsar-Miller M, Azziz R 1998 The development of the polycystic<br />

ovary syndrome: family history as a risk factor. Trends Endocrinol<br />

Metab 9:55–58<br />

489. Franks S, Gharani N, Waterworth D, Batty S, White D, Williamson<br />

R, McCarthy M 1998 Current developments in molecular genetics<br />

of the polycystic ovary syndrome. Trends Endocrinol Metab<br />

9:51–54<br />

490. Taylor JA, Lubahn DB, Impaired glucose tolerance in the ER�KO<br />

mouse. Program of the 80th Annual Meeting of The <strong>Endocrine</strong><br />

Society, New Orleans, LA, 1998 (Abstract P2-15), p 257<br />

491. Mullis PE, Yoshimura N, Kuhlmann B, Lippuner K, Jaeger P,<br />

Harada H 1997 Aromatase deficiency in a female who is compound<br />

heterozygote for two new point mutations in the P450arom gene:<br />

impact of estrogens on hypergonadotropic hypogonadism, multicystic<br />

ovaries, and bone densitometry in childhood. J Clin Endocrinol<br />

Metab 82:1739–1745<br />

492. Sudhir K, Chou TM, Messina LM, Hutchison SJ, Korach KS,<br />

Chatterjee K, Rubanyi GM 1997 Endothelial dysfunction in a man<br />

with disruptive mutation in oestrogen-receptor gene [letter]. Lancet<br />

349:1146–1147<br />

493. Sudhir K, Chou TM, Chatterjee K, Smith EP, Williams TC, Kane<br />

JP, Malloy MJ, Korach KS, Rubanyi GM 1997 Premature coronary<br />

artery disease associated with a disruptive mutation in the estrogen<br />

receptor gene in a man. Circulation 96:3774–3777<br />

494. Rosenfeld CS, Bagegni A, Lubahn DB, <strong>Estrogen</strong> receptor-� knockout<br />

mice reveal a role for estrogen in mammalian female sexual<br />

development. Program of the 80th Annual Meeting of The <strong>Endocrine</strong><br />

Society, New Orleans, LA, 1998 (Abstract OR20–3), p 79<br />

495. Kurita T, Lee KJ, Zhao C, Cooke PS, Taylor JA, Lubahn DB,<br />

Cunha GR 1998 <strong>Estrogen</strong> induces progesterone receptors in uterine<br />

stroma of estrogen receptor-� knockout mouse. Mol Biol Cell<br />

9:238A–238A<br />

496. Kenney NJ, Washburn T, Afshari C, Bocchinfuso W, Korach K,<br />

Barrett JC 1997 The response of null estrogen receptor mammary<br />

parenchyma and mesenchyme in vivo. Mol Biol Cell 8[Suppl]:1976–<br />

1976<br />

497. Young LJ, Wang Z, Donaldson R, Rissman EF 1998 <strong>Estrogen</strong><br />

receptor � is essential in the induction of oxytocin receptor by<br />

estrogen. Neuroreport 9:933–936<br />

498. De Vries GJ, Sannen K, Rissman EF, Steroid-responsive vasopressin<br />

innervation in the brain of estrogen receptor-��mice. Program<br />

of the 27th Annual Meeting of the Society of Neuroscience,<br />

New Orleans, LA, 1997 (Abstract 830.14), p 2139<br />

499. Wersinger SR, Barron KI, Rissman EF 1998 Estradiol decreases<br />

proGnRH immunoreactivity in female wild type but not estrogen<br />

receptor � knock-out mice. Program of the 28th Annual Meeting of<br />

the Society for Neuroscience, Los Angeles, CA, 1998 (Abstract<br />

632.6), p 1607<br />

500. Fugger HN, Rissman EF, Foster TC 1998 <strong>Estrogen</strong> receptor � is not<br />

required for estradiol’s effects on inhibitory avoidance behavior in<br />

female mice. Program of the 28th Annual Meeting of the Society for<br />

Neuroscience, Los Angeles, CA, 1998 (Abstract 84311), p 2121<br />

501. Fugger HN, Schiml PA, Rissman EF, Foster TC, Characterization<br />

of spatial behavior and hippocampal electrophysiology in male<br />

estrogen receptor-� minus mice. Program of the 27th Annual Meeting<br />

of the Society for Neuroscience, New Orleans, LA, 1997, (Abstract<br />

825.10), p 2124<br />

502. Fugger HN, Cunningham SG, Rissman EF, Foster TC 1998 Sex<br />

differences in the activational effect of ER� on spatial learning.<br />

Horm Behav 34:163–170<br />

503. Johns A, Freay AD, Fraser W, Korach KS, Rubanyi GM 1996 Disruption<br />

of estrogen receptor gene prevents 17�-estradiol-induced angiogenesis<br />

in transgenic mice. Endocrinology 137:4511–4513<br />

504. Freay AD, Curtis SW, Korach KS, Rubanyi GM 1997 Mechanism<br />

of vascular smooth muscle relaxation by estrogen in depolarized rat<br />

and mouse aorta. Role of nuclear estrogen receptor and Ca 2� uptake.<br />

Circ Res 81:242–248<br />

505. Cross HR, Steenbergen C, Korach KS, Murphy E 1998 <strong>Estrogen</strong><br />

receptor knock-out mice exhibit higher myocardial contractility<br />

than wild-type mice but are equally susceptible to ischemic injury.<br />

Circulation 98:758–758<br />

506. Roemmich JN, Rogol AD, Hale JE, Lubahn DB, Rissman EF 1997<br />

Bone strain independently augments bone mineral in pubertal mice<br />

with the disrupted estrogen receptor gene. J Bone Miner Res 12:<br />

T563–T563<br />

507. Smithson G, Couse JF, Korach KS, Kincade PW 1997 The role of<br />

estrogen receptors and androgen receptors in sex steroid regulation<br />

of B lymphopoiesis. J Immunol 161:27–34<br />

508. Hurn PD, Sampei K, Sawada M, Goto S, Crain BJ, Alkayed NJ,<br />

Korach KS, Traystman RJ, Demas GE, Nelson RJ, Stroke in mice<br />

deficient in classical estrogen receptors (ERKO�). Program of the<br />

28th Annual Meeting of the Society for Neuroscience, Los Angeles,<br />

CA, 1998 (Abstract 207.2), p 516<br />

509. Couse JF, Lindzey J, Dixon D, Korach KS, Use of the estrogen<br />

receptor-� to investigate the role of ER� in the developmental and<br />

carcinogenic actions of diethylstilbestrol. Program of the 80th Annual<br />

Meeting of The <strong>Endocrine</strong> Society, New Orleans, LA, 1998<br />

(Abstract OR48-2), p 114<br />

510. Yellayi S, Teuscher C, Lubahn DB, Hess RA, Bunick D, Lee KH,<br />

Cooke PS 1998 Role of estrogen receptor-� (ER�) in the development<br />

and function of the thymus in male and female mice. Biol<br />

Reprod 58[Suppl 1]:131–131<br />

511. Dieudonne SC, Xu T, Chou JY, Kuznetsov SA, Satomura K, Mankani<br />

M, Fedarko NS, Smith EP, Robey PG, Young MF 1998 Immortalization<br />

and characterization of bone marrow stromal fibroblasts<br />

from a patient with a loss of function mutation in the estrogen<br />

receptor-� gene. J Bone Miner Res 13:598–608

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!