Successful implantation is the result of reciprocal interactions between the implantation-competent blastocyst and receptive uterus. Although various cellular aspects and molecular pathways of this dialogue have been identified, a comprehensive understanding of the implantation process is still missing. The receptive state of the uterus, which lasts for a limited period, is defined as the time when the uterine environment is conducive to blastocyst acceptance and implantation. A better understanding of the molecular signals that regulate uterine receptivity and implantation competency of the blastocyst is of clinical relevance because unraveling the nature of these signals may lead to strategies to correct implantation failure and improve pregnancy rates. Gene expression studies and genetically engineered mouse models have provided valuable clues to the implantation process with respect to specific growth factors, cytokines, lipid mediators, adhesion molecules, and transcription factors. However, a staggering amount of information from microarray experiments is also being generated at a rapid pace. If properly annotated and explored, this information will expand our knowledge regarding yet-to-be-identified unique, complementary, and/or redundant molecular pathways in implantation. It is hoped that the forthcoming information will generate new ideas and concepts for a process that is essential for maintaining procreation and solving major reproductive health issues in women.

  • I. Introduction

  • II. Preimplantation Embryo Development and Genomic Activation

  • III. Species-Specific Morphological Blueprint and Timing of Implantation

  • IV. Delayed Implantation

  • V. Window of Implantation: A Transient and Unique Moment

  • VI. Embryo-Uterine Signaling Pathways in Implantation

    • A. Steroid hormone signaling

    • B. Signaling via adhesion molecules: cell-cell interactions

    • C. Signaling by vasoactive factors

    • D. Signaling by growth factors

    • E. Signaling by cytokines

    • F. Homeobox genes in implantation

    • G. Ligand-dependent nuclear receptors and coactivators in implantation

    • H. Cell cycle regulation and signaling in implantation and decidualization

    • I. Matrix remodeling and angiogenesis during implantation and decidualization

  • VII. Emerging Concepts

    • A. Endocannabinoid signaling in implantation

    • B. Developmental genes in implantation

    • C. Discovery of novel implantation-related genes

  • VIII. Perspectives and Future Directions

I. Introduction

PROCREATION, BOTH SEXUAL and asexual, is a fundamental evolutionary process necessary to sustain life. Viviparity is a landmark in the process of evolution. Sexual procreation in higher eukaryotes, especially in mammals, is often inferior to asexual procreation in prokaryotes and in some eukaryotes with respect to shear number of progeny. Thus, mammalian reproduction is more complex and highly regulated for the propagation of superior offspring to carry on the task of procreation. The nurturing of an offspring within the body and producing a live birth is an enduring task. This process demands safeguard regulatory systems at various critical steps. The assembly of a new life first depends on the union between a sperm and an egg (ovum) culminating in fertilization; failure to achieve such a union leads to their demise. The one-cell fertilized egg, termed embryo, undergoes several mitotic cell divisions, eventually forming a differentiated tissue called the blastocyst with two distinct cell populations, the inner cell mass (ICM) and a layer of trophectoderm cells surrounding the ICM (1). The embryo proper is derived exclusively from the ICM, whereas the placenta and extraembryonic membranes are produced from cells contributed mainly by the trophectoderm. A two-way interaction between the blastocyst and maternal uterine luminal epithelium initiates the process of implantation, a process by which blood vessels of the embryo are brought into functional communication with the maternal circulation leading to the establishment of a functional placenta and pregnancy. Maternal resources filtered across the selective barrier of the placenta protect and nourish the conceptus. Placental types have been classified into three categories: hemochorial (rodents, humans, and nonhuman primates), epitheliochorial (horses, cows, sheep, and pigs), and endotheliochorial (most carnivores) (reviewed in Ref. 2).

A significant pregnancy loss resulting from preimplantation embryonic death is common to many mammals and is considered to be a selection process leading to the survival of superior embryos for implantation. However, dysregulation of the events before, during, or immediately after implantation also may often be a cause for poor pregnancy rates in eutherian mammals. Understanding the mechanism of preimplantation embryonic development and implantation in the uterus has been a challenge to reproductive and developmental biologists with the goal of alleviating the problems of human infertility and ensuring the birth of quality offspring. Such knowledge is also necessary for developing novel contraceptive approaches to restrict world population growth.

The current state of our knowledge of preimplantation and implantation physiology is the result of the accumulation of scientific observations gathered over many years. Implantation is a complex process involving spatiotemporally regulated endocrine, paracrine, autocrine, and juxtacrine modulators that span cell-cell and cell-matrix interactions. However, the precise sequence and details of the molecular interactions involved have not yet been defined. Furthermore, the implantation process varies among species, thus precluding the formulation of a unified theme. In addition, ethical restrictions and experimental difficulties prevent direct analysis of embryo-uterine interactions during human implantation. Thus, it is an onerous task to write a review on the molecular basis of embryo-uterine interactions during implantation that could be relevant to mammals in general. This review focuses primarily on the physiological and molecular basis of implantation in mice because more mechanistic information is now available for this species. However, an attempt has been made to indicate comparative analyses based on limited work in other species.

Despite experimental success in initiating embryonic development outside the womb and in identifying numerous molecules involved in the embryo-uterine dialogue (37), there is a yet-to-be-filled significant knowledge gap in understanding the in vivo events of implantation. The successful implantation of an embryo is contingent upon cellular and molecular cross-talk between the uterus and the embryo. The coordination of the endocrine, cellular, and molecular events via paracrine, autocrine, and/or juxtacrine factors in a dynamic manner produces within the uterus a favorable environment, the receptive state, to support implantation. The embryo also functions as an active unit with its own molecular program of cell growth and differentiation. Thus, deficiencies in uterine receptivity, embryo development, or the embryo-uterine dialogue will compromise fertility. This review of the implantation process focuses on the molecular basis of embryo homing and attachment, on elucidating the reciprocal signaling networks between the embryo and uterus, and on determining genetic causes of implantation failure.

II. Preimplantation Embryo Development and Genomic Activation

Preimplantation embryo development and differentiation, which culminate in the formation of a blastocyst, require the activation of the embryonic genome, a process that is essential to implantation. The maternal-zygotic transition occurs at the two-cell stage in mice and other rodents, between the eight- and 16-cell stages in cows and sheep, and between the four- and eight-cell stages in humans (reviewed in Ref. 8). Upon activation of the embryonic genome, the embryo grows rapidly to form a blastocyst. At the blastocyst stage, embryos mature and escape from their zona pellucidae to gain implantation competency. The differentiated and expanded blastocyst is composed of three cell types: the outer polarized epithelial trophectoderm, the primitive endoderm, and the pluripotent ICM. The ICM provides the future cell lineages for the embryo proper (9, 10), and the trophectoderm, the very first epithelial cell type in the developmental process, makes the initial physical and physiological connection with the uterine luminal epithelium. The ICM is not identifiable in marsupial blastocysts that appear as a hollow ball of cells with similar morphological characteristics. It is also not known which cells are programmed to form the embryo proper (reviewed in Ref. 10). The formation of the trophectoderm and its subsequent development into trophoblast tissue are crucial steps for the initiation of implantation and the establishment of pregnancy. Trophoblast cells produce a variety of growth factors, cytokines, and hormones that influence the conceptus and maternal physiology in an autocrine, paracrine, and/or juxtacrine manner (11, 12).

Preimplantation embryo development normally occurs within the zona pellucida. However, zona removal by various experimental manipulations does not deter embryonic development in vitro (13), suggesting that this glycoprotein barrier is not essential for development to progress. The nonadhesive nature of the zona pellucida is thought to facilitate the journey of embryos through the oviduct and from the oviduct to the uterus. In mice and rats, normal embryonic development to the blastocyst stage within the reproductive tract requires the presence of ovarian estrogen and progesterone (14). There is a reduced number of embryos and a reduced number of cells per embryo in the absence of these hormones (15), but treatment with estrogen and progesterone reverses these defects (16). Because there is no convincing evidence that estrogen and/or progesterone act directly on the preimplantation embryo (17), embryonic development is considered to depend on growth-promoting factors originating from the reproductive tract under the influence of these hormones. However, apparently normal development in simple defined media in culture suggests that preimplantation embryos are capable of producing their own growth-promoting factors (reviewed in Ref. 18). In fact, several growth factors, cytokines, and their receptors are expressed in the embryo, and the proliferative and differentiating effects of these factors on embryonic development and functions have been observed (reviewed in Refs. 3 and 1820).

III. Species-Specific Morphological Blueprint and Timing of Implantation

Implantation is the process by which the blastocyst comes into intimate physical and physiological contact with the uterine endometrium. Enders and Schlafke (21, 22) have classified the process of implantation into three stages: apposition, adhesion, and penetration. Apposition is the stage when embryonic trophectoderm cells become closely apposed to the uterine luminal epithelium. This is followed by the adhesion stage in which the association of the trophectoderm and the luminal epithelium is sufficiently intimate as to resist dislocation of the blastocyst by flushing the uterine lumen. The stage of penetration involves the invasion of the luminal epithelium by the trophectoderm. Stromal cell differentiation into decidual cells (decidualization) is more extensive, and the loss of the luminal epithelium is evident at this stage. These three stages of implantation form a continuum.

In mammals, especially rodents, a generalized stromal edema occurs before the beginning of apposition. This event leads to the closure of the uterine lumen, which results in interdigitation of the microvilli of the trophectoderm and the luminal epithelia (apposition), followed by closer contact between them (the adhesion or attachment reaction). Luminal closure occurs throughout the entire uterus during pregnancy or pseudopregnancy and thus does not require the presence of blastocysts. Priming of the uterus with progesterone alone appears to be sufficient for this event to occur; luminal closure does not occur in the absence of progesterone. Although luminal closure and apposition occur in progesterone-treated delayed implanting mice, the attachment reaction does not occur. Estrogen treatment is required for attachment to occur.

Bonnet (23), on the basis of different types of blastocyst-uterine cell-cell interactions, classified implantation into three categories: central, eccentric, and interstitial. Central implantation occurs in mammals such as rabbits, ferrets, and some marsupials. In these animals, blastocysts grow and expand extensively before implantation. In contrast, the blastocysts of mice, rats, and hamsters are small and show modest expansion. In these species, an implantation chamber is formed by the invagination of the uterine epithelium, which is a characteristic of eccentric implantation. In guinea pigs, chimpanzees, and humans, the implantation process is of the interstitial type, i.e., blastocysts are embedded within the subepithelial stroma. The results of ultrastructural studies led Schlafke and Enders (24) to classify implantation into intrusive, displacement, and fusion types. In the intrusive type of implantation, which occurs in humans and guinea pigs, the trophoblasts penetrate through the luminal epithelium, reaching and extending through the basal lamina. The displacement type of implantation occurs in rodents; the luminal epithelium is freed from the underlying basal lamina, facilitating the spread of trophoblasts through the epithelium. The fusion type of implantation, in which trophoblasts make a connection with the luminal epithelium by forming symplasma, occurs in the rabbit. In many rodents, including mice and rats, implantation always occurs at the antimesometrial side of the uterus, whereas in some bats, implantation is mesometrial. In other animals, the embryos elongate and either attach over the entire endometrium (horse, pig, and wallaby) or only at specialized areas known as caruncles (cow and sheep) (2). Schematic diagrams for different types of implantation have been illustrated previously (2, 3).

The attachment reaction coincides with a localized increase in stromal vascular permeability at the site of the blastocyst, as can be demonstrated by iv injection of a macromolecular blue dye (uterine blue reaction) (25). The first sign of the attachment reaction (apposition stage) in the process of implantation occurs in the mouse and rat on the evenings of d 4 and d 5, respectively, and on d 6.5 in the rabbit (2527). In the primates, the attachment reaction occurs approximately on d 8 in humans and baboons, on d 9 in macaques, and on d 11 in marmoset monkeys (28, 29). In large domestic animals, the first signs of attachment occur on d 13 in pigs, on d 20 in cows, on d 16 in sheep, and on d 19 in goats (30).

In both mice and humans, stromal cells surrounding the implanting blastocyst undergo decidualization, eventually embedding the embryo into the antimesometrial stromal bed. In mice, blastocysts are oriented with their ICMs directed mesometrially, whereas in humans the ICM is directed antimesometrially. The mechanism by which the blastocyst is directed to the antimesometrial luminal epithelium or by which the orientation of the blastocyst is achieved at the time of implantation remains elusive. There is evidence that in progesterone-treated delayed implanting mice, blastocysts are placed antimesometrially, and interdigitation (apposition) of luminal epithelial cell microvilli occurs with those of the abembryonic or lateral trophectoderm cells of the blastocyst with its ICM oriented toward the uterine lumen. This observation led to the suggestion that upon initiation of the attachment reaction and subsequently the implantation process by estrogen, blastocysts retain the orientation they adopted during delay. During normal implantation in mice with the onset of luminal closure, blastocysts are placed at the antimesometrial side of the lumen along the uterine axis. Shortly after the luminal closure, zona-encased blastocysts are located in implantation chambers with random orientation of the ICMs. However, by the beginning of the attachment reaction, blastocysts are correctly oriented with their ICMs directed at the mesometrial pole. This observation suggested that the trophectoderm of the entire blastocyst surface has the potential for attachment to the luminal epithelium, and that attachment occurs randomly immediately after the loss of the zona pellucida. Evidence was presented to suggest that the correct orientation of the blastocyst is achieved by free movement of the ICM. However, further investigation is necessary to resolve this issue. All of these events, from the luminal closure to the attachment reaction, occur between about 86 and 92 h after coitum in mice (reviewed in Refs. 31 and 32).

IV. Delayed Implantation

Delayed implantation is a process by which implantation is postponed for a certain period. The uterus remains in a quiescent state, and embryos at the blastocyst stage become dormant. Delayed implantation occurs in many vertebrate species, but the underlying mechanisms that direct this process are different in various species that have adapted to this reproductive strategy (33). In mice and rats, ovariectomy before the presumed estrogen surge in the morning of d 4 of pregnancy results in the failure of implantation and initiates a state of dormancy of the blastocyst within the uterine lumen (25, 34). This condition is referred to as delayed implantation and can be maintained for many days by continued treatment with progesterone. The process of implantation with blastocyst activation can be rapidly initiated by a single injection of estrogen in the progesterone-primed uterus (25, 34). The mechanisms by which estrogen mediates the processes of blastocyst activation and implantation are poorly understood. Delayed implantation also occurs naturally (facultative) during lactation after postpartum ovulation and fertilization of the egg in mice and rats (35, 36). However, implantation ensues rapidly after termination of the suckling stimulus. The occurrence of lactational delay in these species is due to the secretion of an insufficient amount of ovarian estrogen. Mustelids, marsupials, and many other species also exhibit obligatory seasonal delayed implantation (3740). Delayed implantation does not occur in some species such as the hamster, guinea pig, rabbit, and pig. Whether this phenomenon occurs in humans is not known. The delayed implantation models in mice and in other species could be exploited more extensively to better understand the molecular signaling that emanates from the embryo and influences uterine biology and vice versa.

V. Window of Implantation: A Transient and Unique Moment

In all eutherian mammals thus far studied, the uterus differentiates into an altered state when blastocysts are capable of effective two-way communication to initiate the process of implantation. This state is termed uterine receptivity for implantation and lasts for a limited period. At this stage, the uterine environment is able to support blastocyst growth, attachment, and the subsequent events of implantation (30, 4143). The major factors that specify uterine receptivity are the ovarian steroids, progesterone and/or estrogens. Ovarian progesterone and estrogen are crucial for implantation in mice and rats, but ovarian estrogen is not essential for implantation in pigs, guinea pigs, rabbits, and hamsters (41, 4448). Estrogen-synthesizing capacity has been demonstrated in rabbit and pig embryos, but whether embryonic estrogen plays a role in implantation in these species is still debatable. Recent evidence suggests that hamster blastocysts express the aromatase gene (Paria, B. C., unpublished results). The mouse embryo lacks aromatase activity necessary for estrogen synthesis (reviewed in Ref. 49). Whether preimplantation estrogen secretion by the ovary or embryo plays a role in human implantation is unknown.

In mice and rats, the coordinated actions of progesterone and estrogen regulating proliferation and/or differentiation of uterine cells in a spatiotemporal manner establish the window of implantation (50). For example, on the first day of pregnancy (vaginal plug) in mice, uterine epithelial cells undergo proliferation under the influence of the preovulatory estrogen secretion. Rising levels of progesterone secreted from freshly formed corpora lutea initiate stromal cell proliferation from d 3 onward. The stromal cell proliferation is further stimulated by a small amount of ovarian estrogen secreted on the morning of d 4 of pregnancy. These coordinated effects of progesterone and estrogen result in the cessation of uterine epithelial cell proliferation, initiating differentiation (50). During normal pregnancy, the presence of an active blastocyst in the uterus is the stimulus for the implantation reaction. After the attachment reaction is initiated on d 4 at 2400 h, stromal cells surrounding the implanting blastocyst begin to proliferate extensively and differentiate into decidual cells (decidualization) (30). In pseudopregnant mice, the steroid hormonal milieu within the uterus is similarly maintained due to the presence of newly formed corpora lutea. Thus, the sensitivity of the pseudopregnant uterus to implantation on d 1–4 is quite similar to normal pregnancy, and blastocyst transfer into the uterine lumen during the receptive phase provokes normal implantation reactions and subsequent decidualization. Although blastocysts are the normal inducers of these events, various nonspecific stimuli, such as intraluminal infusion of oil, air, and mechanical stimuli can also initiate certain aspects of the decidual cell reaction (deciduoma) in pseudopregnant or steroid hormonally prepared uteri (30). However, there is evidence that the initial uterine reactions induced by nonspecific stimuli are different from those induced by blastocysts (51, 52).

Uterine sensitivity with respect to steroid hormonal requirements and implantation has been classified as prereceptive, receptive, and nonreceptive (refractory) phases (30, 41). These phases have been defined by employing embryo transfer experiments in pseudopregnant mice. In the mouse, whereas the uterus is fully receptive on d 4, it is considered prereceptive on d 1–3 of pregnancy or pseudopregnancy. The mouse uterus can be rendered receptive to blastocyst implantation only if exposed to a small amount of estrogen after 24–48 h of progesterone priming (53). Evidence suggests that the uterus is most receptive to implantation on d 4 (43), and the efficiency of implantation decreases with time (54). By d 6, the uterus becomes completely refractory to blastocyst implantation. Recent evidence suggests that concentration of estrogen within a very narrow range determines the duration of the window of uterine receptivity in mice; uterine receptivity remains open for an extended period at lower estrogen levels but rapidly closes at higher levels. Uterine nonreceptivity induced at high estrogen levels is accompanied by aberrant uterine expression of implantation-related genes. These results suggest that careful regulation of estrogen levels could improve female fertility in in vitro fertilization programs (55). Another critical factor determining the window of implantation is the state of activity of the blastocyst, as described below.

In mice and rats, ovariectomy before the preimplantation estrogen secretion on the morning of d 4 of pregnancy induces delayed implantation (34, 43). This status can be maintained for many days if progesterone treatment is continuously provided. Under this condition, blastocysts undergo zona hatching, albeit at a slower pace, but they become dormant without initiating the attachment reaction, and the progesterone-primed uterus remains in the neutral stage. However, a single injection of estrogen promptly induces blastocyst activation with the initiation of implantation in the progesterone-primed uterus. Active and dormant blastocysts are molecularly and physiologically distinguishable. Epidermal growth factor (EGF) receptor (EGF-R), cyclooxygenase-2 (COX-2), and histamine type 2 receptor (H2), the factors that are associated with blastocyst attachment reaction, are expressed in normal or active blastocysts but are down-regulated in dormant blastocysts (43, 5659). In contrast, the G protein-coupled cannabinoid receptor CB1, which is activated by natural and endocannabinoids, is down-regulated in active blastocysts but remains up-regulated in dormant blastocysts (60). Collectively, these findings suggest that a complex array of molecular networks regulates blastocyst activation and dormancy.

Although estrogen is essential for blastocyst activation and implantation in the progesterone-primed mouse uterus, the mechanisms by which estrogen initiates these responses remain elusive. We speculated that estrogen actions in uterine preparation and blastocyst activation for implantation are two distinct events. Embryo transfer experiments in delayed implanting recipient mice provide evidence that whereas the primary estrogen, 17β-estradiol, initiates uterine events for implantation, its catechol metabolite, 4-hydroxy-17β-estradiol (4-OH-E2), participates in activation of dormant blastocysts (17). Blastocyst activation by 4-OH-E2 involves COX-2-derived prostaglandins (PGs) and cAMP (17). The use of an estrogen receptor (ER) antagonist ICI-182,780 showed that, whereas estradiol via its interaction with the nuclear ERs participates in the preparation of the progesterone-primed uterus to the receptive state in an endocrine manner, 4-OH-E2 produced from estradiol in the uterus mediates blastocyst activation in a paracrine manner that does not involve nuclear ERs. These results provide evidence that both primary and catecholestrogens are required for embryo-uterine interactions to ensure successful implantation and that implantation occurs only when uterine receptivity coincides with the activated state of the blastocyst. Molecular pathways that are potentially involved in uterine receptivity and blastocyst activation are discussed below in more detail.

VI. Embryo-Uterine Signaling Pathways in Implantation

Recent advances in molecular and genetic approaches have led to the discovery of numerous molecules involved in embryo-uterine interactions; however, the precise sequence and details of the signaling cascades for many of these molecules have not yet been defined. This review attempts to focus on a select number of signaling molecules and pathways that are implicated in embryo-uterine interactions in relation to implantation (Fig. 1). A large number of other growth factors, cytokines, lipid mediators, and vasoactive agents that could well be involved in implantation are not addressed here due to space limitations.

Fig. 1.

A scheme of signaling networks in embryo-uterine communication during implantation in mice. Implantation in mammals absolutely depends upon synchronized development of the blastocyst to the stage when it is competent to implant and the uterus to the stage when it is receptive to blastocyst growth and implantation. Ovarian estrogen (E2) and progesterone (P4) are the primary effectors that direct the prereceptive uterus to a receptive state via a number of locally expressed growth factors, cytokines, transcription factors, and vasoactive mediators in the uterus, whereas uterine-derived catecholestrogen and regulated levels of endocannabinoids activate the blastocyst to an implantation-competent state. During the attachment phase, signaling and adhesive events embracing the uterus and the blastocyst lead to implantation. Le, Luminal epithelium; Ge, glandular epithelium; S, stroma; CE, catechol estrogen.

A. Steroid hormone signaling

As described earlier, ovarian estrogen is essential for blastocyst implantation in the progesterone-primed uterus in mice and rats. Despite considerable progress regarding the molecular mechanism of estrogen action, many fundamental questions remain unanswered. Estrogen action is normally considered to involve its interaction with a nuclear receptor (ER), a ligand-dependent transcription factor (61, 62). The nuclear ER exists primarily in two isoforms, ERα and ERβ (63, 64). The estrogen-ER complex forms a homodimer that binds to cis-acting estrogen response elements (EREs) in the regulatory regions of the target genes. The EREs for transcription activators are usually present in the 5′-flanking region of specific genes (65). Although a perfect palindromic consensus was identified as AGGTCA(nnn)TGACCT (66), most estrogen-responsive genes have imperfect palindromes or do not have recognizable EREs (65, 6770). Thus, the contention that steroids or their mimics function only by interacting with nuclear receptors to serve as transcription activators for specific DNA response elements is no longer tenable. Several recent reports highlight the complexities in gene regulation by estrogens, including many of the protein-protein interactions mediated via the ER and coregulators (7173). For example, cholera toxin, which has no ER-binding capacity, mimics the mitogenic action of estrogen in the uterus presumably by increasing cAMP levels and protein kinase A activity (74, 75). Protein kinase C can also modulate uterine ER levels, and protein kinase C inhibitors can reduce estrogen-induced mitogenic action (76). These results suggest that membrane-bound receptors acting via protein kinases can increase the expression of the same genes activated by steroid hormone nuclear receptors. This forms the basis for the view that other nonnuclear receptors also interact with steroids or their mimics. Although the presence of a membrane ER was postulated more than two decades ago (77), the subject remains controversial because of the lack of molecular identity. However, signaling by plasma membrane ERs is an emerging concept and requires special attention (78).

A large variety of natural and synthetic compounds that mimic natural estrogens and bind to the nuclear ERα are present in the environment (79, 80). A few of these xenobiotics possess an even higher affinity for ERβ than for ERα (81, 82). The interaction of xenoestrogens with ER is considered as the basis for their reproductive toxicity, although their affinity for the receptor is quite low (83, 84). However, there is evidence that responses to an estrogenic compound in target tissues are not necessarily related to its affinity for the receptor (8588), suggesting the presence of other signaling pathways. Thus, xenoestrogens may interact with ER or other binding proteins that may not result in a similar kind of transactivation that normally occurs with natural ligands. Their estrogen-like effects in the uterus include an increase in water imbibition and ornithine decarboxylase activity, as well as increased DNA and protein synthesis. Xenoestrogens can also induce implantation in progesterone-primed rodents (89). Although the mechanisms of action of these xenoestrogens are not clear, these compounds have received a great deal of attention as a possible cause of certain cancers and impaired reproductive functions (reviewed in Refs. 80 and 90).

Emerging evidence indicates that many of the rapid uterine effects of estrogenic compounds do not involve the classical genomic effects (9195). We and others have observed that ERα mutant female mice, which show negligible classical responses to estrogen (96), nevertheless display estrogenic responses to catecholestrogens or xenoestrogens (9195). For example, early uterine estrogenic responses, including gene expression and macromolecular uptake, are elicited by catecholestrogen or kepone independent of ERα and/or ERβ (91). Furthermore, genetic evidence suggests that several other early responsive target genes are induced in the mouse uterus by primary and catechol estrogens independent of nuclear ERs, suggesting nongenomic downstream events. These downstream signaling pathways represent Wnt signaling, protein processing, and calcium homeostasis (92). In contrast, estrogen, which induces the expression of many growth factors and their receptors and other secretory proteins, may facilitate their transport to the membrane and/or secretion by activating the membrane-trafficking pathway via nuclear ER (97). These results suggest that estrogen regulates diverse but interdependent signaling pathways via ER-dependent and -independent processes.

Despite the evidence for nongenomic steroid hormone actions, differential uterine cell-specific expression of nuclear ERα and ERβ and the progesterone receptor (PR) during the periimplantation period in mice suggests that the coordinated effects of estrogen and progesterone in uterine events for implantation are primarily mediated via these nuclear receptors (98). Furthermore, gene-targeting experiments in mice lacking ER or PR provide valuable information regarding their roles in uterine biology that were not previously recognized. ERα(−/−) mice exhibit infertility due to hyperstimulated ovaries and hypoplastic uteri (99). Furthermore, blastocyst transfers in estrogen- and progesterone-treated uteri of ERα(−/−) mice have established that functional ERα is required for implantation (100). In contrast, ERα(−/−) mice can induce and support decidualization (deciduoma) in response to artificial stimuli if primed appropriately with progesterone alone (101, 102). These results suggest that ERα(−/−) mice fail to initiate and support implantation, perhaps due to the failure of the attachment reaction rather than the failure of decidualization events (91). Indeed, the genes and signaling pathways involved in decidualization are induced in ERα(−/−) after application of an artificial stimulus (100). PR(−/−) mice show pleiotropic reproductive defects including impaired ovulation, uterine hyperplasia, and failure in decidualization (103). Selective deletion of the PR-A isoform also showed infertility but with a milder phenotype, suggesting that PR-A and PR-B serve as functionally distinct mediators of progesterone signaling in vivo (104). The results of experiments that used both PR(−/−) and PR-A(−/−) mice further reinforce a requirement of progesterone in decidualization (103, 104).

Gene-targeting experiments have established the importance of both ER and PR in uterine preparation for implantation in mice (99). However (98, 103), whether the preimplantation embryo is a direct target for steroid hormones remains unclear. Recent evidence suggests that preimplantation estrogen secretion on d 4 of pregnancy in mice has a dual role as primary estrogen and as a catecholestrogen with distinct targets (17). Whereas primary estrogen acts via the uterine ER to prepare the uterus for implantation, catecholestrogens formed locally in the uterus from the primary estrogen participate in blastocyst activation. However, it is still a mystery how catecholestrogens mediate activation of blastocysts (17). Although nuclear ERα is present in both active and dormant blastocysts (105), dormant blastocysts do not respond to estradiol and fail to attain implantation competency in vitro. In contrast, dormant blastocysts do respond to a catecholestrogen 4-OH-E2 and become implantation competent in vitro. The ER antagonist ICI-182,780 fails to reverse this response, suggesting that nuclear ER signaling is not critical to blastocyst activation (17). These observations are surprising in the light of other recent findings that ERα, ERβ, and efp mRNAs are expressed in the preimplantation embryos (106, 107). Examining the direct roles of estrogens and/or progesterone in preimplantation embryo function and how steroid hormone signaling in the embryo and uterus are coordinated for implantation will require further investigation.

Estradiol undergoes hydroxylation to 4-OH-E2 by a P450-linked enzyme CYP1B1 (108). This enzyme is present throughout the mouse uterus on d 4, but disappears from the implantation site on d 5 (17). Activation of dormant blastocysts appears to involve an early response to 4-OH-E2, because dormant blastocysts transferred into delayed implanting recipient uteri within 1 h of estradiol administration of the recipients show implantation, whereas similar blastocysts transferred beyond this 1-h period fail to implant (17). These results suggest that a rapid response that is critical to implantation occurs in utero. In contrast, dormant blastocysts cultured in the presence of 4-OH-E2, but not estradiol, gain implantation competency and, upon transfer, implant in pseudopregnant recipients well beyond the 1-h window of estradiol treatment. Similar results were also obtained by culturing dormant blastocysts in the presence of PGE2 or a permeable analog of cAMP. This effect apparently involves the COX-2 signaling pathway (17). For example, coincubation of dormant blastocysts with a selective COX-2 inhibitor and 4-OH-E2 efficiently blocks their activation and implantation upon transfer to suitable recipients. This effect of the COX-2 inhibitor was partially reversed by addition of PGE2 to the culture media. The results strongly suggest that the action of 4-OH-E2 on dormant blastocysts is mediated via the COX-2 signaling pathway, leading to an increase in intracellular cAMP levels. Further investigation will reveal the types of PGs and receptors involved in this event.

B. Signaling via adhesion molecules: cell-cell interactions

Many glycoproteins and carbohydrate ligands and their receptors are expressed in the uterine luminal epithelium and blastocyst cell surfaces (reviewed in Refs. 109 and 110). Primary adhesion molecules that are implicated in implantation are selectins, galectins, heparan sulfate proteoglycans, Muc-1, integrins, cadherins, and the trophinin-tastin-bystin complex. Muc1 acts as an antiadhesive masking molecule (111). Muc1, a stretch of long carbohydrate moieties, is expressed in the mouse uterine epithelium before implantation. The physical hindrance created by these branches is thought to prevent interaction between the embryo and the luminal epithelium of the uterus before the attachment reaction. This is consistent with timely down-regulation of Muc1 from the luminal epithelium throughout the uterus before the attachment reaction in mice (reviewed in Ref. 3). In contrast, overall Muc1 expression increases in the rabbit and human uterus during the receptive period. However, careful examination revealed that there is indeed a decrease in Muc1 levels at the site of implantation in rabbits (112). In humans, the situation appears to be more complicated. During the apposition phase, the presence of an embryo increases the levels of Muc1 in the epithelium, but at the adhesion phase, the embryo induces a cleavage of Muc1 at the implantation site (113). Collectively, these findings suggest that Muc1 acts as an antiadhesive molecule that must be removed from the site of implantation.

Among the adhesion molecules, integrins have been studied more extensively in the human endometrium because of their cycle-dependent changes and the potential role in uterine receptivity. The members of the integrin family serve as receptors for various extracellular matrix (ECM) ligands and modulate cell-cell adhesion and signal transduction events (114). Each integrin is comprised of two subunits, α and β, and each αβ-combination has its own binding specificity and signaling properties. As membrane-associated receptors, integrins possess short cytoplasmic tails with no enzymatic activity. Signaling by integrins is mediated by associating adaptor proteins that bridge them to the cytoskeleton, cytoplasmic kinases, and transmembrane growth factor receptors (114). Several members of the integrin family, including αvβ3, are known to interact with the RGD (Arg-Gly-Asp) peptide sequence present in many ECM proteins, such as fibronectin, laminin, and entactin. Although many integrin heterodimers show constitutive expression in the uterine epithelium or stroma, α1β1, α3β1, α6β1, αvβ3, and αvβ1 heterodimers exhibit cycle-dependent changes (115119). Of special interest is the expression of α1β1, which shows implantation-related changes. For example, the expression of α1β1 is restricted to early- and midsecretory phases both in the epithelium and stroma and primarily restricted to the stroma during the predecidual phase (117). Furthermore, unexplained infertility in women is associated with the deficiency of αvβ3 or α4β1 in the uterus during the window of implantation (120). However, the functional definition of these markers in uterine receptivity or under pathological conditions still awaits further investigation. In mice, αvβ3 is expressed in both the uterine luminal epithelium and the blastocyst during implantation. It has been shown that an intrauterine injection of RGD peptide or neutralizing antibody against αvβ3 reduces the number of implantation sites in mice and rabbits (121).

Among many subunits, α5β1, α6β1, and αvβ3 are expressed in the mouse embryo throughout the periimplantation period, whereas several others exhibit stage-specific expression (122). Integrins are also expressed in the differentiating trophoblasts at later stages (122), suggesting their roles in trophoblast differentiation and adhesion. A role for fibronectin via integrin binding in blastocyst outgrowth was further confirmed in vitro using antibodies against the αv, α5, β1, or β3, which inhibited adhesiveness on the outer surface of the trophoblast inducible by fibronectin (123). In addition, a gene-targeting experiment revealed that deletion of the β1 gene results in ICM defects and embryonic lethality (124). However, the mutant embryos form morphologically normal blastocysts and initiate implantation, but trophoblast invasion becomes defective (124). Adhesion-competent, late-blastocyst-stage trophoblasts undergo intracellular signaling initiated upon ligation of α5β1 and αvβ3 by fibronectin (125). Integrin signaling mobilizes cytoplasmic Ca2+ and induces the trafficking of intracellular vesicles, resulting in stronger adhesion to fibronectin at the apical surface. Therefore, blastocyst adhesion to the endometrium during implantation is considered to be regulated by the endogenous developmental program, as well as through interactions with ECM components in the local environment (126, 127). Although there is evidence that the embryo is a site of action for integrin signaling, it is not yet clear whether the uterus is a site of action. Results of gene-targeting experiments of integrin subunits are not very informative in relation to their roles in implantation because of the complex phenotypes and apparent compensation by other subunits (128131).

Invasive mouse trophoblasts adhere, spread, and migrate on ECM substrates (132135) and penetrate three-dimensional ECM structures (136, 137). Several ECM components that are up-regulated in the periimplantation endometrium, including fibronectin, laminin, and collagen type IV (138140), support trophoblast outgrowth in vitro (132, 133). Trophoblast interactions with the ECM are mediated primarily by integrins (122, 123, 133135, 141). Hexapeptides containing the RGD sequence recognized by integrins (142) block trophoblast outgrowth on fibronectin, collagen type II and IV, entactin, and vitronectin (133, 135, 141). However, trophoblast adhesion to type I laminin is independent of its RGD sequence and is primarily mediated through interaction of α7β1 with the E8 integrin-recognition domain of laminin (143, 144).

Trophinin was identified by cDNA library screening of a human trophoblastic cell line (145). This transmembrane protein can mediate homophilic interactions between two different cell types. For example, it mediates an interaction between a human endometrial cell line and a trophoblastic cell line, but this interaction is complex (145). Trophinin requires the presence of a cytoplasmic protein tastin to sustain adhesion between these two cell types. In addition, the presence of bystin, another cytoplasmic protein, is required for effective interaction between trophinin and tastin. This adhesion complex, which is present in both trophoblastic teratocarcinomas and endometrial adenocarcinomas, mediates adhesion between them. In humans and monkeys, trophinin is specifically expressed in cells involved in implantation. Furthermore, the trophinin complex was detected in both trophoblast and decidual cells at the human fetalmaternal interface as early as the sixth week of pregnancy (146). Although trophinin expression in mice coincides with the timing of implantation, it is expressed in the luminal and glandular epithelium throughout the uterus irrespective of the presence or absence of blastocysts, raising doubts regarding its role specific to implantation (147).

E-cadherin, a calcium-dependent cell-cell adhesion molecule, participates in the formation of the epithelial adherens junctions in cooperation with α- and β-catenins (148, 149). E-cadherin is a critical factor for blastocyst formation, because its targeted deletion leads to defective embryonic development resulting in failure to form the trophectoderm (150, 151). E-cadherin is implicated in uterine-embryo interactions because of its homotypic adhesive activity (152). Whereas the trophectoderm highly expresses E-cadherin, the components of the adherens junctional complex are also expressed in the uterine luminal epithelium at the time of the attachment reaction. The expression subsequently becomes evident in the subepithelial stroma surrounding the implanting blastocysts with apoptosis occurring in the luminal epithelium (152, 153). Therefore, it is speculated that temporal and cell-specific expression of the adherence junction proteins in the uterus results in molecular guidance, which is important for blastocyst attachment and subsequent invasion.

A recent study has highlighted the critical role of the selectin adhesion system in human implantation (154). This adhesion system, which is involved in leukocyte capture from the bloodstream, is also operative during implantation and placentation. On the maternal side, selectin oligosaccharide ligands are expressed in the receptive uterine epithelium, and on the embryonic side, trophoblast cells expressed L-selectin receptors. It was observed that beads coated with specific selectin ligands adhere to the trophoblast, suggesting that the trophoblast cell surface receptors are functional. This investigation suggests that trophoblast L-selectin mediates interactions with the uterus to establish an adhesion mechanism for implantation. This is an exciting finding in several respects. First, this study affirms that specific ligand-receptor signaling pathways between the embryo and uterus are critical for implantation and subsequent pregnancy establishment. Second, it shows that the same mechanism operative during implantation is also operative during the later phases of pregnancy. Third, it shows that trophoblast cells share a system known to be active in the blood-vascular system. Similar adhesion signaling between the transmembrane form of heparin-binding EFG-like growth factor (HB-EFG) expressed on the luminal epithelial surface and ErbB receptors present on the trophectoderm cell surface for implantation in mice and humans has been reported previously (57, 155).

C. Signaling by vasoactive factors

It has long been speculated that vasoactive agents, such as histamine and PGs, are involved in many aspects of reproduction including ovulation, fertilization, implantation, and decidualization. Histamine functions as a ubiquitous mediator of cell-cell signaling and is synthesized from L-histidine by histidine decarboxylase (HDC) both in peripheral tissues and in the nervous system (156). Histamine is a well-known neurotransmitter in the brain (157), but it is also involved in other physiological responses including gastric acid secretion, regulation of allergic reactions, and vascular permeability (reviewed in Ref. 158). Because the process of implantation is considered a proinflammatory reaction and because increased vascular permeability at the site of blastocyst implantation is common to many species, it was suggested that histamine plays a role in implantation and decidualization (158). Earlier observations suggested that uterine mast cells are a possible source of histamine, and its release from mast cells by estrogen is important for implantation (159, 160). This suggestion was based on the observations that local histamine application stimulates uterine hyperemic and edematous responses (161) and that there is a reduction in uterine mast cell numbers and histamine content after estrogen treatment and during implantation (162). Furthermore, a histamine antagonist pyrathiazine or an inhibitor of HDC was shown to interfere with implantation when instilled into the uterine lumens of rats and rabbits (159, 163). Histamine works via at least four histamine receptor subtypes (H1, H2, H3, and H4) (164166) and blocking both H1 and H2 receptors was shown to interfere with implantation (167). Subsequent studies also showed that histamine induces implantation in delayed implanting rats when injected with a suboptimal dose of estrogen (168). However, successful implantation and birth of live offspring in mast cell-deficient mice and other evidence suggest that uterine mast cell histamine is not essential for implantation (136, 169, 170). Thus, if histamine is involved in implantation, it should be provided either by major uterine cell types or by embryonic cells. A recent study showed that mouse blastocysts do not have the capacity for histamine synthesis (59). However, HDC is expressed in uterine epithelial cells on d 4 of pregnancy in mice before implantation, but not in decidual cells (171). Thus, although histamine may have a role in implantation in mice, its role in decidualization is unlikely. Whereas H1, H2, and H3 receptor subtypes are not detectable in the uterus, H2 receptors are expressed in preimplantation mouse blastocysts. These observations, as well as the inhibition of blastocyst zona-hatching and implantation by H2 antagonists and an HDC inhibitor, suggest that uterine histamine targets the blastocyst for implantation (59). However, apparently normal implantation occurs in mice lacking HDC or H2 type histamine receptor genes, suggesting the possible involvement of other vasoactive agents with overlapping functions in this process (165, 172).

PGs possess vasoactive, mitogenic, and differentiating properties (173) and are implicated in various female reproductive functions. COX, which exists in two isoforms, COX-1 and COX-2, is the rate-limiting enzyme in the biosynthesis of PGs. COX mediates the conversion of arachidonic acid into PGH2, which is then converted to various PGs by specific synthases (173). The COX isoforms are encoded by two separate genes and exhibit distinct cell-specific expression, regulation, and subcellular localization, yet share similar structural and kinetic properties. COX-1 is considered to be a constitutive enzyme that mediates housekeeping functions. In contrast, COX-2 is an inducible enzyme and is induced in a variety of cell types by growth factors, cytokines, and inflammatory stimuli (173). PGs normally exert their function by interacting with cell surface G protein-coupled receptors, but they can also function as ligands for nuclear peroxisome proliferator-activated receptors (PPARs) (174177). Because COX-2 is primarily responsible for increased PG production during inflammation, this isoform is the target for development of selective antiinflammatory drugs (178, 179). COX-2 overexpression is also associated with tumorigenesis (180, 181).

The processes of ovulation and implantation are considered analogous to proinflammatory responses, and thus participation of PGs in these events has been speculated (182, 183). For example, PGs are considered to participate in follicular rupture during ovulation (reviewed in Ref. 184). This is consistent with gonadotropin-mediated induction of COX-2 in ovarian follicles preceding ovulation (85, 184). PGs are also implicated as important mediators of increased endometrial vascular permeability during implantation and decidualization (185). A unique pattern of expression of Cox-1 and Cox-2 genes in the periimplantation mouse uterus further suggests that PGs play important roles in these processes (185). Cox-1 is expressed in uterine luminal and glandular epithelial cells on the morning of d 4 of pregnancy, but its expression becomes undetectable in the luminal epithelial cells by the time of the attachment reaction. In contrast, Cox-2 is expressed in the luminal epithelium and underlying stromal cells solely at the site of blastocyst attachment. Using the delayed implantation model, this study also established that the expression of Cox-2 in the receptive uterus requires the presence of active blastocysts. The results suggested that Cox-2 expression during the attachment reaction is critical to implantation (185). Indeed, gene-targeting experiments have demonstrated that COX-2-derived PGs are essential for implantation and decidualization (186188). Experiments with Cox-1(−/−) mice suggest that the loss of COX-1 is compensated by the expression of Cox-2 for implantation (189). Among various PGs, the levels of prostacyclin (PGI2) are highest at the implantation sites of wild-type mice, and implantation defects are partially restored in Cox-2(−/−) mice by administration of a more stable PGI2 agonist, carbaprostacyclin (190).

The role of PGs is further illustrated by the reduced fertility of female mice lacking cytoplasmic phospholipase A2, which is involved in the liberation of arachidonic acid from membrane phospholipids for PG synthesis by the COX system (191193). The reduced fertility in these females is due to deferral of on-time implantation, leading to subsequent retarded fetoplacental development and reduced litter size (54). Collectively, these results indicate that the cytoplasmic phospholipase A2-COX-2 axis is crucial to implantation. However, one recent study described that wild-type (B6C3H) blastocysts transferred into COX-2-deficient female mice on a mixed (C57Bl6/JX129S7/SvEvBrd) background on d 3 of pseudopregnancy implanted and produced live offspring, although decidual growth was retarded (194). Because this study did not follow experimental protocols similar to other studies, the discrepancy between this study and others cannot be compared scientifically. However, some of the results reported by Cheng and Stewart (194) remain uninterpretable. For example, dissected decidual weights reported by these investigators seem abnormally high on various days of pregnancy, and the sizes of d 7 implantation sites shown appeared larger than d 8 implantation sites and at a more advanced stage of pregnancy. Nonetheless, recent evidence demonstrates that Cox-2 is expressed either in the uterus, blastocyst, or both during implantation in a variety of species with different modes of implantation, including sheep, mink, skunk, baboon, and pig (195198). COX-2 expression in human endometrium has also been reported (199, 200). These results suggest a conserved function of COX-2 in implantation in various species. However, it has been discovered recently that, depending on the genetic background, COX-1 can rescue female infertility in COX-2-deficient mice (201).

PGs exert diverse functions using both cell surface PG receptors and PPARs. Receptors for PGE2, PGF, PGD2, PGI2, and thromboxanes have been named as EP1–EP4, FP, DP, IP, and TP, respectively; they belong to the G protein-coupled family of cell surface receptors (reviewed in Refs. 202 and 203). Although PGE2 synthase is expressed at the implantation sites with the presence of PGE2 and EP receptors (203205) and although PGE2 has been shown to be associated with implantation and decidualization (206), gene-targeting experiments show that three of the four EP receptor subtypes (EP1–EP3) are not critical for implantation. EP4 deficiency results most frequently in perinatal lethality, and thus its role in implantation has not yet been determined (reviewed in Ref. 203). Furthermore, mice deficient in FP or IP show normal implantation. PGs can also exert their effects by utilizing PPARs that belong to a nuclear hormone receptor superfamily. The evidence that PG-mediated PPAR signaling is involved in implantation is discussed below in more detail. PGs also appear to be important for embryonic functions relevant to preimplantation embryo development and implantation. Preimplantation embryos produce PGs, and inhibitors of PG synthesis have been shown to inhibit embryonic growth, functions, and zona hatching in vitro (207 , 208). Dormant mouse blastocysts can achieve implantation competence if cultured in the presence of PGE2 or a permeable analog of cAMP. This effect apparently involves the COX-2 signaling pathway (17). However, normal development of Cox-1(−/−) /Cox-2(−/−) double-mutant embryos in the uterus suggests that PGs of embryonic origin are not essential for embryo development (208). However, compensation by maternal PGs in embryonic development cannot be ruled out. Vascular endothelial growth factor (VEGF), also known as vascular permeability factor, is highly vasoactive in nature. It is a potent inducer of vasodilation and angiogenesis. Its role in implantation is discussed below.

D. Signaling by growth factors

The expression of various growth factors and their receptors in the uterus in a temporal and cell-specific manner during the periimplantation period suggests that these factors are important for implantation (36, 209, 210). The present review highlights primarily the importance of the EGF family of growth factors; however, the roles of other growth factors, such as TGFβs, fibroblast growth factors (FGFs), IGFs, platelet-derived growth factors, and many others should not be ignored. The EGF family of growth factors includes EGF itself, TGFα, HB-EGF, amphiregulin, betacellulin, epiregulin, and neuregulins (58, 211). HB-EGF is the earliest molecular marker found in the uterus exclusively at the sites of active blastocysts appearing several hours before the attachment reaction in mice (26). This induction is followed by the expression of betacellulin, epiregulin, neuregulin-1, and Cox-2 around the time of the attachment reaction (58, 185, 211). In contrast, amphiregulin is expressed throughout the uterine epithelium on the morning of d 4 of pregnancy and is well characterized as a progesterone-responsive gene in the uterus (212). Around the time of the attachment reaction, strong expression of amphiregulin in the luminal epithelium is found only around the implanting blastocysts, and this expression is absent by the morning of d 5. Although these results suggested that amphiregulin has a role in implantation, amphiregulin-deficient mice or compound knockout mice for EGF/TGFα/amphiregulin do not exhibit implantation defects (213, 214). Because HB-EGF, betacellulin, epiregulin, neuregulin, and amphiregulin all show overlapping uterine expression patterns around the implanting blastocyst at the time of attachment reaction (reviewed in Refs. 3 and 211), it is assumed that a compensatory mechanism rescues implantation in the absence of one or more members of the EGF family.

The EGF-like growth factors interact with the receptor subtypes of the erbB gene family, which is comprised of four receptor tyrosine kinases: ErbB1 (EGF-R), ErbB2, ErbB3, and ErbB4. They share common structural features but differ in their ligand specificity and kinase activity (215). The initial dimerization between coexpressed receptors upon ligand binding constitutes the classical mechanism of action of EGF-like ligands. Spatiotemporal expression patterns of EGF gene family members and ErbBs in the uterus during the periimplantation period suggest compartmentalized functions of EGF-like growth factors in implantation (58).

A number of growth factors and their receptors are expressed in preimplantation embryos of several species, suggesting their roles in preimplantation mammalian development (216, 217). In this review, we focus on potential roles of the EGF family of ligands with respect to preimplantation embryo development and implantation. ErbB1 (EGF-R), ErbB2, and ErbB4, the receptor subtypes for the EGF family of growth factors, are expressed in the mouse blastocyst (Refs. 56 and 218 and our unpublished results), and EGF or TGFα has beneficial effects on embryonic development in vitro (18). Using genetic and biochemical approaches, the roles of embryonic ErbB1 and/or ErbB4 in interacting with uterine HB-EGF in blastocyst implantation have recently been highlighted in mice (57, 218). HB-EGF is expressed as soluble and transmembrane forms in the uterine luminal epithelium at the site of the blastocyst before the attachment reaction, suggesting paracrine and/or juxtacrine interactions with embryonic ErbBs, as well as autocrine, paracrine, and/or juxtacrine interactions with uterine ErbBs that are expressed in a spatiotemporal manner during the periimplantation period (26, 58, 219). For example, the expression of both ErbB1 and ErbB4 is down-regulated in dormant blastocysts during delayed implantation but is readily up-regulated with blastocyst activation and initiation of implantation (56, 218). Furthermore, whereas a recombinant soluble HB-EGF can promote blastocyst growth and differentiation (26), cells that express the transmembrane form of HB-EGF can adhere to active, but not dormant, blastocysts in vitro (57), suggesting paracrine and juxtacrine functions of HB-EGF. In addition, by directing an HB-EGF-toxin conjugate toward wild-type and erbB1(−/−) blastocysts, it was shown that HB-EGF could also interact with embryonic ErbB4 and heparan sulfate proteoglycan molecules (218). Collectively, these results suggest that an interaction between uterine HB-EGF and blastocyst ErbBs is important for the attachment reaction. However, the absolute necessity of HB-EGF in implantation requires genetic evidence. A recent report shows that most HB-EGF mutant mice die early in postnatal life due to cardiac defects, precluding critical examination of the implantation phenotype (220). It is also to be noted that early events of implantation do not appear to be affected by blastocysts deficient in either ErbB1 or ErbB4 (221, 222), although the implantation-initiating efficiency of blastocysts deficient in more than one receptor type needs to be tested to delineate the functional redundancy among the receptor family. In conclusion, detailed expression and gene-targeting experiments with all of the ligands and receptors are required to define paracrine, autocrine, and/or juxtacrine roles of specific ligand or its receptors in implantation.

Among many growth factors that have been studied in humans, HB-EGF appears to play a role in implantation and embryonic development. Its expression is maximal during the late secretory phase (cycle d 20–24) when the endometrium becomes receptive for implantation (223, 224) and cells expressing the transmembrane form of HB-EGF adhere to human blastocysts displaying cell surface ErbB4 (155). Furthermore, HB-EGF was shown to be one of the most potent growth factors for enhancing the development of human in vitro fertilization-derived embryos to blastocysts and subsequent zona hatching (225). Thus, cumulative evidence suggests that HB-EGF has a significant role in preimplantation embryo development and implantation as a paracrine and/or juxtacrine factor in various species.

E. Signaling by cytokines

The expression of various cytokines and their receptors in the uterus and embryo during early pregnancy suggests their roles in various aspects of implantation (reviewed in Refs. 3 , 20 , 225 , and 226). However, gene-targeting studies show that mice lacking TNFα, IL-1β, IL-1 receptor antagonist, IL-1 receptor type 1, IL-6, and granulocyte/macrophage-colony stimulating factor apparently do not manifest overt reproductive defects (reviewed in Ref. 20). These observations suggest that either these molecules have minor roles in implantation or the loss of one cytokine is compensated by other cytokines with overlapping functions. In contrast, some cytokines are important for normal female fertility (227229). For example, female op/op mice with a naturally occurring mutation of the M-CSF gene have markedly impaired fertility (227), and mice with a null mutation of the Lif gene encoding leukemia inhibitory factor (LIF) show complete failure of implantation, and blastocysts in these mutant mice undergo dormancy (228, 230). Studies using IL-11Rα mutant mice have also shown that IL-11 is crucial to decidualization, but not for the attachment reaction (229). Interestingly, both LIF and IL-11 are members of the IL-6 family, which includes IL-6 itself, oncostatin M, ciliary neurotrophic factor, and cardiotrophin (231). LIF and IL-11 bind to ligand-specific receptors, LIFR and IL-11R, respectively, and share gp130 as a signal transduction partner (231), suggesting that gp130 signaling is critically involved in implantation. Although the mechanism underlying implantation and decidualization failures in the absence of LIF still remains to be elucidated, recent evidence shows that there is a loss or an aberrant expression of certain implantation-related genes in pregnant Lif mutant mice (232). For example, uterine expression of HB-EGF and epiregulin is absent, and Cox-2 expression is aberrant at the sites of blastocysts in Lif mutant mice during the anticipated time of implantation.

LIF and its receptors, LIFR and gp130, exist in both soluble and membrane-bound forms, and soluble forms of these two receptors antagonize the actions of their ligands, implying the complexity of the LIF signaling pathway (233235). Lif is transiently expressed in uterine glands on d 4 of pregnancy in mice, suggesting its role in implantation (236). However, our recent studies show that uterine Lif expression is biphasic on d 4. Not only is Lif expressed in glands, but it is also expressed in stromal cells surrounding the blastocyst at the time of the attachment reaction (232). This suggests that LIF has dual roles: first in the preparation of the uterus, and later in the attachment reaction. However, the molecular mechanism by which LIF executes its effects on implantation is not yet known. In this regard, it would be useful to establish the complicated ligand-receptor interactions and detailed expression patterns of LIF receptors that occur during the periimplantation period. However, a recent report shows that inactivation of gp130 by deleting all signal transducers and activators of transcription binding (STAT) sites results in implantation failure (237), reinforcing the importance of LIF signaling in implantation.

The uterine milieu in Lif mutant mice fails to induce implantation irrespective of the blastocyst genotypes, since Lif(−/−) blastocysts can implant after transfer to wild-type pseudopregnant recipients (228, 230). These reciprocal embryo transfer experiments suggest that maternal LIF is essential for blastocyst implantation. However, a role for this cytokine in embryonic functions cannot be ignored, because LifR and gp130 are expressed at the blastocyst stage, and administration of exogenous LIF improves embryo viability and hatching in several species (238240). Taken together, these data suggest that both the preimplantation embryo and the uterus are sites of LIF action. However, embryos lacking either LIFR or gp130 develop to the blastocyst stage and implant normally but die during the perinatal period (241, 242). These results raise questions about the role of LIF signaling in preimplantation embryo development.

Lif expression in the uterus is maximal around the time of implantation in most species examined, although the steroid hormonal requirements for the preparation of uterine receptivity and implantation differ depending on the species. Whereas uterine Lif expression in several species appears to be regulated by P4 (reviewed in Ref. 242), estrogen regulates Lif expression in the mouse uterus. This is evident from Lif expression on d 1 of pregnancy and during the estrous stage of the cycle when the uterus is under the influence of estrogen stimulation (236, 243, 244). In addition, Lif is not expressed in the uterus during experimentally induced delayed implantation but is rapidly induced by an injection of estrogen (232, 236). However, it has yet to be learned how estrogen induces Lif expression in the mouse uterus and the mechanism by which it is regulated by progesterone in other species. In humans, Lif is expressed in the endometrium and at higher levels in the glandular epithelium of the secretory endometrium (245). Furthermore, LIF deficiency has been associated with unexplained recurrent abortions and infertility in women (246).

F. Homeobox genes in implantation

Hox genes are transcription factors that belong to a multigene family. They are developmentally regulated and share a common highly conserved sequence element called the homeobox that encodes a 61-amino acid helix-turn-helix DNA-binding domain (247). Hox genes are organized in four clusters (A, B, C, and D) on four different chromosomes in mice and humans and follow a stringent pattern of spatial and temporal colinearity during embryogenesis (247). Several Hox genes at the 5′-end of each cluster are classified as AbdB-like Hox genes, because of their homology with the Drosophila AbdB gene. In vertebrates, AbdB-like Hox genes, similar to their Drosophila ortholog, are expressed in developing genitourinary systems (248). For example, Hoxa-10 and Hoxa-11 are highly expressed in developing genitourinary tracts and the adult female reproductive tract, suggesting roles in reproductive events (248250). Hoxa-10 mutant mice exhibit oviductal transformation of the proximal one third of the uterus. Furthermore, adult female mice deficient in Hoxa-10 show failures in blastocyst implantation and decidualization unrelated to the oviductal transformation (248). Subsequent studies revealed that uterine stromal cells in Hoxa-10-deficient female mice show reduced proliferation in response to progesterone, leading to decidualization defects (248, 251). The uterus in Hoxa-11-deficient mice is hypoplastic and devoid of uterine glands due to developmental defects. Defective proliferation of stromal cells in Hoxa-10(−/−) female mice suggests that Hoxa-10 is involved in the local events of cellular proliferation by regulating cell cycle molecules. Indeed, cyclin D3 is aberrantly expressed in Hoxa-10 mutant uteri in response to a decidualizing stimulus (252). Furthermore, because several progesterone-responsive genes are dysregulated in the uterine stroma of Hoxa-10 mutant mice (251), Hoxa-10, as a transcriptional factor, may convey progesterone responsiveness in the uterine stroma by regulating gene expression. A similar, but more severe, phenotype was also noted in Hoxa-11-deficient female mice (250).

A recent study using microarray analysis further revealed that the absence of Hoxa-10 in the uterus in response to progesterone is associated with two cellular disturbances (253). First, among many genes that were up-regulated in the Hoxa-10-deficient uteri, two cell cycle molecules, p15 and p57, were notable. These two genes are both cyclin-dependent kinase inhibitors (CKIs), suggesting that the previously observed defect in stromal cell proliferation in Hoxa-10 mutant mice could be associated with this up-regulation of CKIs. Second, the microarray experiments and follow-up fluorescence activated cell sorter analyses demonstrated that there was hyperproliferation of T lymphocytes in the Hoxa-10-deficient uterine stroma in response to progesterone. These results suggest that an aberrant lymphoproliferation has adverse effects on implantation in the Hoxa-10-deficient mice. In humans, both Hoxa-10 and Hoxa-11 genes are markedly up-regulated in the uterus during the midsecretory phase in steroid hormone-dependent manner (254), suggesting their roles in human implantation.

There is a homeobox gene family, unrelated to other larger classes of homeobox genes, called the Hmx family of transcription factors. These genes show overlapping expression during development, but gene-targeting experiments have revealed a unique role for Hmx3 in female reproduction (255). Hmx3 mutant female mice show normal fertilization and preimplantation embryo development to blastocysts. However, the blastocysts fail to implant in the uterus and subsequently die. Because Hmx3 is primarily expressed in the myometrium during early pregnancy, the mechanism of infertility in these mice is different from that of Hoxa-10 or Hoxa-11 mutant mice and has yet to be explored.

G. Ligand-dependent nuclear receptors and coactivators in implantation

The nuclear receptor superfamily of transcription factors modulates expression of target genes by binding to specific DNA elements. The members of this superfamily span from well-characterized steroid hormone receptors to orphan nuclear receptors with no known ligands. Steroid hormone receptors aside, the PPAR family of nuclear receptors has been implicated in female reproductive events. Three members of the PPAR family are PPARα, PPARγ, and PPARδ. To act as a transcriptional activator, PPARs must form a heterodimer with a member of the retinoid X receptor (RXR) subfamily (176, 190). As described below, whereas developmental defects are common in PPARδ and PPARγ mutant mice, no apparent reproductive phenotype is evident in PPARα mutant mice.

1. PPARδ(β)

PGs can act via dual receptor signaling systems, either via classical cell surface receptors or through nuclear receptor systems. PPARs can respond to a wide variety of ligands including natural and synthetic eicosanoids, fatty acids, and hypolipidemic and hypoglycemic drugs (256). There is evidence that COX-2-derived PGI2 participates in implantation via activation of PPARδ (190), because the implantation defects in Cox-2(−/−) mice are reversible by a PGI2 agonist or a combination of PPARδ and RXR agonists. PGI2 is the most abundant PG in the early pregnant mouse uterus and is higher at implantation sites than in interimplantation sites (190). Consistent with the finding that COX-2-driven uterine PG production is crucial to implantation, Cox-2 and prostacyclin synthase (Pgis) are coexpressed at the implantation site, suggesting the availability of PGI2 directly to uterine cells. In searching for a target receptor for PGI2 in the uterus, the expression of known PGI2 receptors, such as IP, PPARα, and PPARδ, was examined. Among these, PPARδ was colocalized at similar regions of the implantation sites with Cox-2 and Pgis; the expression of IP and PPARα was very low to undetectable. The functionality of PPARδ as a PGI2 receptor was further examined in vivo, using COX-2-deficient mice as an in vivo model. Administration of cPGI or L-165041 (a selective PPARδ agonist) to these mice improved implantation and decidualization (190). In conjunction with other in vitro evidence, this work suggests that PPARδ expressed in the uterine stroma responds to a PGI2 agonist to mediate embryo implantation (190, 257, 258). Three independent groups have reported diverse phenotypes of PPARδ knockout mice (259261). Because of severe early developmental defects of PPARδ mutant embryos, it is very difficult to utilize this model to directly address whether the absence of maternal PPARδ affects implantation as in COX-2-deficient mice. Therefore, the conditional knockout mouse model with uterine-specific deletion of PPARδ is necessary to address this issue.

2. PPARγ

PPARγ is well known as a key metabolic regulator participating in obesity control, diabetes, atherosclerosis, and other processes (262). PPARγ-activating ligands include naturally occurring compounds, such as 15-deoxy-Δ (12, 14)-PGJ2, and the thiazolidinedione class of insulin-sensitizing synthetic compounds (262). Whereas the roles of PPARγ in several metabolic processes have been well documented, gene-targeting experiments reveal that PPARγ is also required for normal development of placental, cardiac, and adipose tissues. PPARγ mutant mice die around embryonic d 10 due to defective terminal differentiation of the trophoblast and placental vascularization (263). Furthermore, in a study using a tissue-specific PPARγ knockout mouse model generated by cross-breeding of mouse mammary tumor virus-cAMP response element and conditional PPARγ-null mice, it was shown that deletion of ovarian PPARγ in female mice led to impaired implantation (264). However, the cause of uterine dysfunction is not due to the absence of uterine PPARγ, because PPARγ was not deleted in the uterus by this cross-breeding. Further study is required to elucidate the underlying mechanism of impaired uterine function in the absence of ovarian PPARγ.

3. Transcriptional cofactors

Many transcription factors including nuclear receptors modulate transcription by direct binding to sequence-specific DNA response elements in promoters of target genes, resulting in activation or repression of transcription in a promoter-specific manner (265, 266). Although biological functions and profiles of activating ligands are extremely diverse, ligand-activated nuclear receptors utilize a converging point of transcriptional cofactors to activate/repress downstream target genes (267). Thus, although many cofactors were cloned based on their interaction with a specific nuclear receptor, various studies show a wide range of sharing of cofactors among the family of nuclear receptors (268), supporting the notion of a functional convergence in vivo. Because of the promiscuous nature of transcriptional cofactors, they have been extensively studied in relation to various nuclear receptors. Recently, several gene-targeted mouse models demonstrate that many of these cofactors are involved in developmental and reproductive processes.

a. cAMP response element binding protein (CREB)-binding protein (CBP)/p300

CBP and the related p300 are called cointegrators because of their many-sided interactions with nuclear receptors, cofactors, and basal transcriptional machinery (268). Thus, as a common limiting cofactor for diverse transcriptional activators and coactivators, CBP apparently organizes multiple signals into an integrated response at promoters containing multiple cis-acting elements (268). CBP generally exhibits constant levels of expression in various cell lines and in the developing embryo (269). Furthermore, the gene dosage-dependent role of CBP and p300 is well demonstrated in experiments with knockout mice in that CBP(−/−), p300(−/−), and CBP(+/−)xp300(+/−) double heterozygous mice all die in utero during midgestation due to multiple developmental defects (269). This suggests an essential role of cointegrators in embryonic development.

b. Steroid receptor coactivator (SRC)

The SRC family consists of three members: SRC-1, transcriptional intermediary factor 1 (TIF1) (SRC-2/glucocorticoid receptor interacting protein 1), and p/CIP(SRC-3/AlB1/RAC3/ACTR). The SRC family members interact with a variety of nuclear receptors including ER, PR, and PPAR, indicating their roles in various reproductive functions (270, 271). The three SRC family members share similar properties with respect to interacting with nuclear receptors and transcriptional activity, but it is suggested that these factors exhibit diverse expression patterns and functions (268). Whereas SRC-1-deficient mice are viable and fertile, they show reduced hormone responsiveness in several target organs, including the uterus, prostate, and testis (272). In contrast, fertility of both male and female TIF1(−/−) mice is impaired (273). Female TIF1(−/−) mice exhibit increases in embryonic resorption between d 12.5 and 18.5 gestation, possibly resulting from placental hypoplasia. The placental hypoplasia in TIF1(−/−) female mice seems to be due to the maternal deficiency of TIF1 in decidual stromal cells. Lastly, SRC-3(−/−) mice also exhibit reduced fertility primarily due to decreased ovulation (274). These results suggest that three members of SRC family cofactors play diverse roles in reproductive processes.

c. PPAR-binding protein (PBP)

PBP was first identified in a yeast two-hybrid screening system as a PPARγ cofactor. This 165-kDa protein was also independently cloned by other groups as vitamin D receptor-interacting protein 205 or thyroid hormone receptor (TR)-associated protein 220, each being a part of a multiunit complex for transcriptional activation by vitamin D receptor or TR, respectively (275, 276). PBP is widely expressed in various mouse tissues and can interact with ER and PPAR (277, 278). Gene-targeting experiments showed that PBP(−/−) embryos die in utero during midgestation due to poor placental development (279, 280).

d. PPAR-interacting protein (PRIP)/nuclear reaceptor-activating protein 250

PRIP or nuclear receptor-activating protein 250 was identified as a PPARγ or PPARα cofactor independently (281, 282). PRIP interacts with members of the PPAR, retinoic acid receptor and RXR families, as well as ER and TR. This gene is widely expressed in various tissues including reproductive organs (281, 282), and PRIP(−/−) mice exhibit severe developmental defects showing placental and cardiac hypoplasia (283, 284).

e. Receptor-interacting protein 140 (RIP140)

Two independent groups cloned RIP140 as a cofactor for PPARα or TR2 (an orphan receptor) (285, 286). This factor actively competes with SRC-1 in transcriptional assays and therefore works as a true transcriptional corepressor (285). RIP140 is involved in female fertility, as demonstrated in gene-targeting experiments (287). RIP140 mutant female mice show reduced ovulation that leads to small litter size, and a follow-up study showed that uterine function in RIP140-deficient mice is normal (288).

H. Cell cycle regulation and signaling in implantation and decidualization

As stated above, the differentiation of the uterus to support embryo development and implantation is primarily directed by progesterone and estrogen (25, 43). Decidualization is first initiated at the antimesometrial site where blastocysts implant. This process, characterized by stromal cell proliferation and differentiation into specialized type of cells (decidual cells) with polyploidy, is critical to the establishment of pregnancy in many species. The mechanisms by which the cell cycle events govern decidualization are poorly understood. The cell cycle is tightly regulated at two checkpoints, the G1-S and G2-M phases. Normal operation of these phases involves a complex interplay of cyclins, cyclin-dependent kinases (cdks), and CKIs. At the beginning of decidualization, the stromal cells immediately surrounding the implanting blastocyst proliferate (50). In mice, stromal cells close to the embryo cease to proliferate, initiating the formation of the primary decidual zone (PDZ) later on d 5 of pregnancy; the PDZ is fully established by d 6. However, stromal cell proliferation outside the PDZ continues, eventually forming the secondary decidual zone (SDZ) (30). Under normal conditions, the stimulus for decidualization is the implanting blastocyst. However, a similar process (deciduoma) can be experimentally induced in the pseudopregnant or hormonally prepared rodent uterus by intraluminal infusion of various agents including oil (30). The development of decidua or deciduoma in rodents is associated with the formation of multinucleate and giant cells (289291). In mice, the decidual cells in the antimesometrial zone are characterized by polyploidy and endoreduplication, and most cells in this zone eventually enlarge, containing nuclei with as much as 64n DNA.

The well-known regulators of mammalian cell proliferation are the three D-type cyclins (D1, D2, and D3), also known as G1 cyclins (292). The D-type cyclins accumulate during the G1 phase. Their association with cdk4 or cdk6 is important for forming holoenzymes that facilitate cell entry into the S phase. The retinoblastoma protein (Rb) and its family members, p107 and p130, are negative regulators of the D-type cyclins. Inactivation of these regulators by phosphorylation is dependent on the cyclin/cdk complex activity and allows the cell cycle to progress through the G1 phase (293). Overexpression of D-type cyclins shortens the G1 phase and allows rapid entry into S phase (294). In contrast, cyclins A and B are involved in the progression from the S through G2-M phase. Binding of cyclin A or cyclin B to cdk1 induces phosphorylation and activation of the complex that is essential to the G2-M phase transition, whereas the cyclin A/cdk2 complex participates during progression in the S phase. In general, the action of cdks is constrained by at least two families of CKI, p16 and p21. The p16 family includes p15, p16, p18, and p19, and they inhibit the catalytic partners of D-type cyclins, cdk4 and cdk6. The p21 family consists of p21, p27, and p57, and they inhibit cdks with a broader specificity. CKIs accumulate in quiescent cells, but are down-regulated with the onset of proliferation. Thus, a critical balance between the positive and negative cell cycle regulators is a key decision maker for cell division (295).

The uterus is a unique and dynamic physiological model in which cellular proliferation, differentiation, polyploidization, and apoptosis occur in a spatiotemporal manner during the reproductive cycle and pregnancy. In the human uterus, various cyclins (A, B1, D1, and E) and cyclin-dependent kinases (cdk1, cdk2, and cdk4) and CKI (p27) are regulated during the menstrual cycle or after hormone treatments (57, 58). These molecules are expressed primarily in epithelial and stromal cells during the proliferative phase, suggesting their involvement in rhythmic proliferation of these cells (296). During the secretory phase or after progesterone administration, p27 expression correlates with progesterone-induced growth suppression in endometrial glands and stromal basalis (297). In rodents, however, the uterine expression of D- and E-type cyclins and cdks is regulated by estrogen and/or progesterone in a temporal manner (298300). Progesterone-dependent growth suppression of the endometrium is considered to be mediated by decreased cdk activity, which presumably occurs via decreased levels of cyclins and increased association of CKI (p27) with cdks (301). There is also evidence that progesterone inhibition of uterine epithelial cell proliferation is mediated by the inhibition of nuclear translocation of cyclin D1 and cdk4 in association with the activation of cyclins A- and E-dependent cdk2 activity (302). Cell cycle molecules are also involved in endoreduplication in trophoblast differentiation during placentation (303, 304).

In mice, the expression of cyclin D3 is up-regulated in decidualizing stromal cells at the implantation site and is associated with cell proliferation (252) (305, 306). Furthermore, cyclin D3 is associated with the large polyploid cells that are defined as terminally differentiated stroma. A schematic model, as shown in Fig. 2, illustrates a possible role for cyclin D3 together with other cell cycle molecules in the developmental regulation of stromal cell decidualization and polyploidy. Cyclin D3 generally associates with cdk4 and/or cdk6 for cell proliferation. The coordinate expression of cdk4 and cyclin D3 at the site of the embryo after the onset of implantation in mice on d 5 of pregnancy suggests that these regulators play roles in proliferation of stromal cells undergoing decidualization. However, the expression of p21 with concomitant down-regulation of cyclin D3 and cdk4 in the PDZ at the implantation site in the afternoon of d 5 supports the view that cell proliferation activity of cdk4/cyclin D3 ceases with the development of the PDZ. Their expression in the decidualizing stroma outside the PDZ is again consistent with their role in proliferation of the stroma at the SDZ. In contrast, down-regulation of cdk4 in the SDZ with persistent expression of p21 on d 6 of pregnancy perhaps directs differentiation of stromal cells in this zone. On this day of pregnancy, a switch from cdk4 to cdk6 with continued expression of cyclin D3 and p21 in stromal cells within the SDZ is noted in polyploid decidual cells. The presence of cyclin E, cyclin A, and cdk2 with concomitant down-regulation of cyclin B and cdk1 in these cells supports the view that these cells are following the endocycle pathway.

Fig. 2.

A proposed model of stromal cell polyploidy and decidualization. The phase-specific cell cycle regulators in the G1 phase (cyclin D3, p21, cdk4, cdk6, cyclin E, and cdk2) or in the S-G2-M phases (cyclin A, cyclin B, cdk1, and cdk2) are shown with respect to their association in mitotic cell cycle vs. endocycle. The coordinate regulation of cyclin D3 and cdk4 in decidualizing stromal cells suggests that these regulators play roles in proliferation. However, the expression of p21 with concomitant down-regulation of cyclin D3 and cdk4 supports the view that cell proliferation ceases with the development of the PDZ. Furthermore, a switch from cdk4 to cdk6 with sustained expression of cyclin D3 and p21 in cells within the SDZ is consistent with the progression through the G phase for the onset of endocycle. This is further supported by the expression of cyclin E, cyclin A, and cdk2 in the polyploid cells for successful progression in G and S phases of the endocycle pathway. The absence of cyclin B and cdk1 probably plays a role to initiate the first endocycle.

The physiological significance of stromal cell polyploidy during decidualization is still unclear. The life span of decidual cells during pregnancy is limited, and their demise makes room for the rapidly growing embryo. Because most decidual cells become polyploid during their lifetime, it is speculated that polyploidy limits the life span of decidual cells. Furthermore, one of many uterine functions is to support embryonic growth that requires increased protein synthesis by decidual cells. The developing polyploidy may ensure increased synthetic capacity by increasing the number of gene copies for transcription. In conclusion, a tight coordination of the cell cycle molecules appears to be critical for uterine cell proliferation and differentiation during implantation and decidualization.

I. Matrix remodeling and angiogenesis during implantation and decidualization

Tissue remodeling and angiogenesis are two hallmark events during implantation and decidualization. The changing endocrine state of the female during the reproductive cycle and pregnancy results in extensive remodeling in the uterine tissue (139, 307). For example, various basement membrane components, such as type IV collagen, laminin, fibronectin, and proteoglycans, in the human uterus undergo changes throughout the menstrual cycle and pregnancy (307). Likewise, the ECM components undergo remodeling during mouse uterine stromal cell decidualization (139). Matrix metalloproteinases (MMPs) and tissue inhibitors of MMPs (TIMPs) are thought to be key mediators for matrix degradation during implantation and decidualization (reviewed in Refs. 307309). There is evidence that a balance between a select set of MMPs and TIMPs is important for implantation. Mechanisms regulating the MMP and TIMP genes during the periimplantation period are not clear, although growth factors and cytokines including the EGF and TGFβ family members and LIF have been shown to modulate MMPs and TIMPs (reviewed in Ref. 309).

Under physiological conditions, angiogenesis, the process by which new blood vessels develop from preexisting vessels, primarily occurs in the uterus and ovary of the adult during the reproductive cycle and pregnancy (310). Indeed, increased vascular permeability and angiogenesis are crucial to successful implantation, decidualization, and placentation (Fig. 3). A number of studies provided indirect and descriptive evidence for the potential roles of estrogen and progesterone in these processes in various species (reviewed in Refs. 310312). These studies primarily examined the changes in the whole uterus of the expression of a number of gene products known to regulate vascular permeability and angiogenesis, including VEGF and its receptors, without investigating its angiogenic status. Thus, in vivo roles for estrogen and progesterone in uterine angiogenesis are not fully appreciated. VEGF, originally discovered as a vascular permeability factor (reviewed in Ref. 313), is also a potent mitogen for endothelial cells and a key regulatory growth factor for vasculogenesis and angiogenesis (314). Targeted disruption of even one allele of the Vegf gene results in embryonic death in utero during midgestation with aberrant blood vessel formation (315, 316). Differential splicing of the Vegf gene generates several VEGF isoforms in both humans and mice; VEGF121 and VEGF165 are the predominant isoforms in humans, whereas VEGF120 and VEGF164 are the most abundant isoforms in mice (311, 317).

Fig. 3.

A scheme showing angiogenic signaling in the uterus during implantation. Increased vascular permeability and angiogenesis at the site of the blastocyst are two hallmarks of the implantation process. The proangiogenic factor VEGF and its receptor Flt1 (VEGFR1) and Flk1 (VEGFR2) are primarily important for uterine vascular permeability and angiogenesis before and during the attachment phase of the implantation process, whereas VEGF in complementation with the angiopoietins (Ang1 and Ang2) and their receptor Tie-2 directs angiogenesis during decidualization. Furthermore, HIFs and COX-2-derived prostaglandins PGs are important for uterine angiogenesis during implantation and decidualization and primarily target the VEGF, but not the angiopoietin, system. Ang1 in collaboration with VEGF induces vessel maturation and maintains vessel leakiness, whereas Ang2 induces vessel destabilization required for further sprouting in the presence of VEGF. Flk1lacZ mice were used to study angiogenesis during implantation. LacZ-stained (blue) blood vessels are shown at the implantation site on d 8 of pregnancy. EC, Endothelial cells; NO, nitric oxide; bFGF, basic FGF; IFN-α, interferon α.

VEGF effects are primarily mediated by two tyrosine kinase receptors: VEGFR1 [fms-like tyrosine kinase 1(FLT1)] and VEGFR2 [fetal liver kinase 1 (FLK1)/kinase insert domain-containing receptor (KDR)] (318321). Although FLT1 activation does not stimulate endothelial cell mitosis, targeted disruption of the Flt1 gene produces impaired endothelial cell assembly into blood vessels and embryonic lethality (322). FLK1 is the major transducer of VEGF signals that induce chemotaxis, actin reorganization, and proliferation of endothelial cells (314, 323, 324). Targeted deletion of the Flk1 gene in mice produces defects in hematopoietic and endothelial cell development leading to embryonic death by d 9.5 (325). Recently, another multifunctional VEGF receptor has been identified as Neuropilin-1 (NRP1). NRP1 was originally described as a neuronal transmembrane receptor that participates in axonal guidance in the developing nervous system (326, 327) and is a receptor for the collapsin/semaphorin family of proteins (328, 329). It is now known that NRP1 functions as a receptor for at least five different ligands, collapsin-1/semaphorin-IIII/D, semaphorin-E, semaphorin-IV, VEGF165, and placental growth factor (PlGF), which are involved in different biological processes, such as nervous system development, vasculogenesis, and angiogenesis (329, 330). NRP1 is expressed in human endothelial cells as a VEGF165-specific receptor. When coexpressed in endothelial cells with FLK1, NRP1 enhances the binding of VEGF165 to FLK1 and VEGF165-mediated chemotaxis severalfold higher than that of FLK1 alone (331). Conversely, inhibition of VEGF165 binding to NRP1 inhibits its binding to FLK1 and its mitogenic activity in endothelial cells. NRP1-deficient mice show peripheral nervous system abnormalities and die in midgestation due to yolk sac vascular insufficiency and developmental anomalies of the cardiovascular system (332). Mice overexpressing NRP1 also show cardiovascular abnormalities including increased number of blood vessels and abnormal hearts (333). Several structurally similar VEGF family members including PlGF, VEGF-B, VEGF-C, and other VEGF-related proteins have recently been identified, and they are distinct gene products. VEGF-C and PlGF can bind to FLK1 and FLT1, respectively. VEGF-C can also interact with a third structurally related tyrosine kinase receptor, FLT4 (reviewed in Ref. 312). Definitive physiological significance of these newly identified VEGF-related growth factors in uterine angiogenesis during implantation warrants further investigation.

We have recently shown that the genes encoding murine VEGF isoforms and their receptors, FLT1, FLK1, and NRP1, are differentially expressed in the mouse uterus in a spatiotemporal manner during implantation and that the predominant VEGF164 isoform interacts with FLK1 and NRP1 (311, 312). These results suggest that the VEGF system is important for uterine vascular permeability and angiogenesis during implantation. Others have also shown the expression of VEGF and its receptors in the uterus as a whole during pregnancy and in response to steroid hormones (313). For example, estrogen rapidly induces uterine vascular permeability and Vegf expression transcriptionally via nuclear ER (313), and the Vegf gene contains EREs (334). Progesterone also up-regulates uterine Vegf expression via activation of the nuclear PR but at a slower rate (334). Because estrogen rapidly stimulates uterine vascular permeability and Vegf expression, and because vascular permeability is considered a prerequisite for angiogenesis, it is widely believed that estrogen is a potent stimulator of uterine angiogenesis during normal reproductive processes in vivo. However, recent evidence using molecular, genetic, physiological, and pharmacological approaches shows that estrogen and progesterone have different effects in vivo; estrogen promotes uterine vascular permeability but profoundly inhibits angiogenesis, whereas progesterone stimulates angiogenesis with little effect on vascular permeability. These effects of estrogen and progesterone are mediated by differential spatiotemporal expression of proangiogenic factors in the uterus (335).

VEGF effects are complemented and coordinated by another class of angiogenic factors, the angiopoietins (336). VEGF acts during the early stages of vessel development (315, 316, 325), whereas angiopoietin-1 (Ang1) acts later to promote angiogenic remodeling, including vessel maturation, stabilization, and leakiness (337339). In contrast to agonistic functions of Ang1, Ang2 behaves as an antagonist. Thus, Ang1 and Ang2 are naturally occurring positive and negative regulators of angiogenesis, respectively. They interact with an endothelial cell-specific tyrosine kinase receptor Tie2 (340). Two additional members of the angiopoietin family have been identified recently. Ang3, which is expressed in mice, appears to function as an antagonist to Ang1 activation of Tie2 in a fashion similar to Ang2 (341). In contrast, Ang4, the human counterpart of Ang3, functions as an agonist to Tie2 (341). However, definitive biological functions of Ang3 and Ang4 remain unclear. Our recent investigation shows that whereas VEGF and its receptor Flk1 are primarily important for uterine vascular permeability and angiogenesis before and during the attachment phase of the implantation process, VEGF together with the angiopoietins and their receptor Tie2 directs angiogenesis during decidualization after implantation (342). A recent report has shown that Ang2 is required for postnatal angiogenic remodeling, and Ang2 in collaboration with VEGF participates in the development of lymphatic vasculature (343).

PGs, because of their roles in angiogenesis, cell proliferation, and differentiation in other systems, are also likely to participate in uterine vascular permeability and angiogenesis during implantation and decidualization. Indeed, there is now genetic and molecular evidence that COX-2-derived PGs participate in uterine angiogenesis during implantation and decidualization (342). Thus, one cause of failure of implantation and decidualization in Cox-2(−/−) mice is the deregulated vascular events in the absence of COX-2. The attenuation of uterine angiogenesis in these mice is primarily due to defective VEGF signaling rather than the angiopoietin system. Collectively, the results provide evidence that whereas ovarian steroid hormones influence uterine vascular permeability and angiogenesis during the preimplantation period, COX-2-derived PGs direct these events during implantation and decidualization by differentially regulating VEGF and angiopoietin signaling (335, 342).

Oxygen homeostasis is essential for cell survival and is primarily mediated by hypoxia-inducible factors (HIFs). These factors are intimately associated with vascular events and induce Vegf expression by binding to the hypoxia response element in the Vegf promoter. HIFα isoforms function by forming heterodimers with the aryl hydrocarbon nuclear translocator (ARNT) (HIF-β) family members. There is very limited information on the relationship among HIFs, ARNTs, and VEGF in the uterus during early pregnancy, although roles of HIFs in regulating VEGF and angiogenesis in cancers and vasculogenesis during embryogenesis are well documented (344). Using molecular and physiological approaches, we have recently shown that uterine expression of HIFs and ARNTs does not correlate with Vegf expression during the preimplantation period (d 1–4) in mice. In contrast, their expression follows the localization of uterine Vegf expression with increasing angiogenesis during the postimplantation period (d 5–8). This disparate pattern of uterine HIFs, ARNTs, and Vegf expression on d 1–4 of pregnancy suggests that HIFs have multiple roles in addition to the regulation of angiogenesis during the periimplantation period. Using pharmacological, molecular, and genetic approaches, we also observed a novel finding that whereas progesterone primarily up-regulates uterine HIF-1α expression, estrogen transiently stimulates that of HIF-2α (345). The definitive role of hypoxia in uterine angiogenesis warrants further investigation.

VII. Emerging Concepts

A. Endocannabinoid signaling in implantation

Psychoactive cannabinoids are active components of marijuana that work via activation of the G protein-coupled cell surface receptors, CB1 and CB2 (346, 347). The discovery of cannabinoid receptors led to the identification of endogenous cannabinoid ligands, arachidonoylethanolamide (anandamide) and 2-arachidonoylglycerol (2-AG) (348, 349). The mouse uterus synthesizes anandamide, and the levels fluctuate in the uterus during early pregnancy coincident with the window of uterine receptivity for implantation (350). Thus, anandamide levels were found to be lower in the receptive uterus and at the implantation sites, but were higher in the nonreceptive uterus and at interimplantation sites (350). The progesterone-treated delayed implanting uterus also showed elevated levels of anandamide, but the levels were down-regulated with the termination of the delayed implantation by estrogen (60). Experiments using Lif(−/−) pregnant mice also supported this finding, because these mice with implantation failure had higher uterine levels of anandamide than those of wild-type mice (60). Therefore, a correlation between the levels of anandamide and phases of uterine receptivity suggests that endocannabinoid ligand-receptor signaling is an important aspect of defining the window of uterine receptivity for implantation (350). There is experimental evidence that ligand-receptor signaling with cannabinoids directs preimplantation embryo development and implantation. This is consistent with the observation that mouse blastocysts express CB1 (351). CB1 is expressed from the two-cell stage at the time of zygotic gene expression through the blastocyst stage (351). Embryonic CB1 is functional, because two-cell embryos cultured in the presence of natural, synthetic, and endocannabinoids fail to develop to the blastocyst stage, and this failure occurs between the eight-cell and blastocyst stages. This effect is reversed by a CB1-selective antagonist (352). Furthermore, the endocannabinoid anandamide at a low concentration stimulates blastocyst differentiation and trophoblast outgrowth, whereas at higher concentrations it inhibits these events via differential regulation of MAPK and Ca2+ signaling (353, 354). These results suggest that a narrow range of cannabinoid concentrations regulate the embryonic developmental program.

We speculate that a tightly regulated level of uterine anandamide and embryonic CB1 during early pregnancy is important for preimplantation embryonic development and implantation. This speculation is consistent with our recent observations of asynchronous preimplantation embryo development in CB mutant mice, and inhibition of implantation in wild-type mice, but not in CB1(−/−) × CB2(−/−) double-mutant mice, with experimentally induced sustained levels of exogenously administered cannabinoids (60). Collectively, the expression of cannabinoid receptors in the preimplantation mouse embryo, synthesis of anandamide in the uterus, and the dose- and stage-specific effects of anandamide on embryo development and implantation suggest that ligand-receptor signaling with endocannabinoids or cannabinoid agonists is important for these events. The observation that heightened levels of cannabinoids inhibit implantation in mice subsequently led to the discovery that elevated levels of anandamide induce spontaneous pregnancy losses in women (355, 356). Thus, regulated cannabinoid signaling perhaps functions as a physiological surveillance system that assures implantation of healthy, but not abnormal, embryos resulting from aberrant expression of CB1 or their exposure to aberrant levels of endogenous or exogenous cannabinoids (Fig. 4). The present observation identifies unique biological functions of cannabinoid signaling for a very fundamental process that determines the survival or demise of a developing embryo.

Fig. 4.

A schematic diagram showing potential endocannabinoid signaling in blastocyst activation and implantation. Anandamide and 2-AG are the major endocannabinoids that interact with G protein-coupled cannabinoid receptors, CB1 and CB2. Regulated levels of endocannabinoids in the receptive uterus and CB1 in activated blastocysts at the time of implantation are beneficial for implantation, whereas higher levels are detrimental to this process. Because COX-2 is expressed in the uterus at the site of implantation and because anandamide and 2-AG can serve as substrates for either COX-2 or FAAH, the proposal suggests that uterine endocannabinoids are tightly regulated by the coordinated activity of FAAH and COX-2 in the uterus during early pregnancy. Evidence suggests that regulated uterine levels of endocannabinoids and blastocyst CB1 play a physiological role in synchronizing blastocyst competency with uterine receptivity for implantation. Tr, Trophectoderm; ANA, anandamide; CB1, brain-type cannabinoid receptor; FAAH, fatty acid amide hydrolase; IS, implantation site; Inter-IS, interimplantation site.

B. Developmental genes in implantation

The intricate cross-talk between the blastocyst and uterus during implantation has several features of the reciprocal epithelial-mesenchymal interactions during embryogenesis and involves evolutionarily conserved signaling pathways. An emerging concept is that many of these evolutionary conserved genes, including those encoding FGFs, IGFs, bone morphogenetic proteins (BMPs), Wnts, Noggin, Indian hedgehog (IHH) proteins, and their receptors, are potential players in the process of implantation and embryo spacing in the uterus (52) (Fig. 1). These genes are expressed in the uterus in a spatiotemporal manner during the periimplantation period in mice. For example, the attachment reaction is associated with a localized stromal induction of genes encoding BMP-2, FGF-2, and WNT-4. A simple in vitro model of implantation is not yet available to examine either the hierarchy of the events elicited in the uterus by the embryo or the function of individual signaling proteins. These questions were addressed by selectively delivering factors via blastocyst-sized gelatin beads in the uterine lumen to provoke implantation-like reactions with correct gene expression similar to what is generated by living embryos during normal implantation. We observed that beads soaked in HB-EGF or IGF-I, but not other proteins, induce many of the same discrete local responses elicited by the blastocyst, including increased localized vascular permeability, decidualization, and expression of Bmp-2 and COX-2 at the sites of the beads (4, 52). Furthermore, beads containing BMP-2 do not themselves produce an implantation-like response but alter the spacing of implantation sites induced by blastocysts cotransferred with the beads (52).

Genes encoding the components of the hedgehog-signaling pathway, namely IHH, the multipass transmembrane HH-binding protein/receptor, PATCHED (PTC), and the transcription factors, GLI 1, -2, and -3 (357359) are expressed in a dynamic temporal and spatial pattern during the preparation of the uterus for implantation (360). The expression of Ihh increases in the luminal epithelium and glands from d 3, reaching high levels on d 4. During the same time, the expression of Ptc, Gli1, and Gli2 is up-regulated in the underlying mesenchymal stroma. Transcription of Ihh in ovariectomized mice is induced by progesterone but not by estrogen. Lower induction of Ihh, Ptc, and Hoxa-10 is also observed in response to progesterone treatment in the uteri of PR mutant mice lacking the progesterone nuclear steroid receptor. This finding suggests that this hormone regulates Ihh via both nuclear receptor-dependent and -independent pathways. Furthermore, in uterine explant cultures, recombinant N-SHH (Sonic hedgehog with N-terminal acylation) protein stimulates the proliferation of mesenchymal cells and the expression of noggin. These findings suggest that IHH generated by the epithelium functions as a paracrine growth factor for stromal cells during the early stages of pregnancy (360). Progesterone regulation of Ihh expression in the mouse uterus has recently been confirmed by another group of investigators (361). Research involving the genes known to be involved in developmental processes in the field of implantation biology is in its infancy and needs to be pursued more vigorously.

C. Discovery of novel implantation-related genes

The identification and characterization of molecules involved in embryo implantation have relied historically on candidate gene approaches with a requirement for knowledge of gene structure or sequence. Exceptions to this include the use of subtractive hybridization, representational difference analysis, and similar techniques (362366). These approaches have identified a number of implantation-related genes in uteri of rodents and primates including humans, but these studies did not provide information on the expression profile of large gene families. Another approach is to modify unknown genes by insertion of transgenic markers or by chemical inactivation. These large-scale mutagenesis screens offer the ability to randomly inactivate genes throughout the genome and detect sought-after phenotypic changes, such as infertility. Although this approach does not address the expression of multiple genes, it does lead to recognition of important molecules without prior information on gene identity or structure. Unfortunately, there are no mutagenic studies that have identified genes implicated in implantation. Differential display (DDRT-PCR) and serial analysis of gene expression (SAGE) are RT-PCR-based methods that do not require advance knowledge of genetic sequence information. Although there are no published reports that used the SAGE technique to identify implantation-related genes, expression studies using DDRT-PCR have identified a small number of interesting genes that are differentially expressed in the uterus during the periimplantation period (252, 362, 367372). The DDRT-PCR technique identified a large number of differentially expressed RT-PCR products that are candidates for further evaluation. To date, none of the study designs using DDRT-PCR to examine the implantation process overlap, so the comparison of their results shows little similarities.

The advent of cDNA microarray technology and sequencing of the mouse and human genomes have led to approaches for global analysis of implantation-related genes. Gene array and SAGE results are comparable (373, 374). Neither method requires prior knowledge of gene sequence, and both provide a read-out on the expression profile of a significant segment of the genome at any given time point or under any experimental condition. A series of uterine gene array studies have produced a large number of implantation-specific candidate genes. In animal models, gene arrays have been used to identify genes that show differential expression at implantation vs. interimplantation sites (375), implantation vs. postimplantation periods (376), or genes that are differentially regulated by changes in estrogen and progesterone signaling (253, 361, 377381). We compared the results of implantation vs. interimplantation arrays to the results of progesterone-treated (delayed implantation) uteri vs. estrogen-stimulated (termination of delayed implantation) uteri to find genes that may be important for implantation. The combined strategy identified a number of genes with previously recognized roles in implantation, confirming the validity of this approach. Among genes that had decreased expression at the implantation site and during progesterone-induced delay, we found a marked shift in the expression of immune-related genes, suggesting active modification of the host immune response to the implanting blastocyst at an early stage. A recent microarray analysis examining the uterine response to progesterone identified a similar subset of Ig-family genes (253). This study identified 1675 progesterone-induced genes that were grouped into 18 unique patterns of progesterone responsiveness. About one half (750 genes) of these genes were also detected by our combined array analyses. Other microarray studies have used PR-deficient mice (361) and RU-486-treated mice (377) to identify novel progesterone-mediated pathways during implantation. After comparing and analyzing the results available for each of these disparate approaches, we selected a subset of genes that are commonly expressed in the mouse uterus under varying periimplantation-like conditions (Table 1). Closer examination of these genes and comparison of results from similar microarray experiments may provide clues to the identity of important genes or gene families in the implantation process.

TABLE 1.

Identification of implantation-related genes in the mouse

Progesterone-responsiveEstrogen-responsive
Up-regulatedDown-regulated
Alkaline phosphataseArachidonate 15-lipoxygenaseBM-90/fibulin
AmphiregulinATFxEIG 180 (ethanol-induced gene)
Apg-2 (chaperone)Cytosolic adenylate kinaseGlutathione S-transferase, θ-2
Carbonic anhydrase IIGADD45 proteinHereditary hemochromatosis-like protein
Cathepsin FGlutamyl-tRNA synthetaseHoxd4
CCAAT/enhancer binding protein βGuanine nucleotide regulatory proteinHS1-binding protein 3
Chondroitin sulfate proteoglycan 2Heat shock protein, 105 kDaIntracisternal A particles
Claudin-7Hexokinase II, exon 1Leptin receptor
Complement C1q B chainIL-1 receptor, type 11Norrie disease homolog
Cyclin-dependent kinase inhibitor 1CMitochondrial stress-70 proteinP glycoprotein 3
Dickkopf-3NAD-dependent methylenetetrahydrofolate dimethyl cyclohydrolaseRas-related protein (DEXRAS1)
FollistatinThioether S-methyltransferase
Glutathione-S-transferaseNM23 metastatic-associated proteinXeroderma pigmentosum, complementation group C
Histidine decarboxylaseNuclear autoantigenic sperm protein
Hoxa 11p45 MAPK kinase
IGF binding protein-3Procollagen, type VI, α 2
IL-13 receptor, 2Protein kinase inhibitor p58
Keratin complex 1RAB geranylgeranyl transferase
LactotransferrinRAMP3
Leukocyte 12/15 lipoxygenaseRan GTPase
LRG-21RAN GTPase-activating protein 1
Membrane metalloendopeptidaseRNAse L inhibitor (Mu-RLI)
Metallothionein 1Small proline-rich protein 2F
Norrie disease homologSplicing factor, arginine/serine-rich 10
Osteoblast-specific factor 2Squalene epoxidase
Peptidylarginine deiminaseSqualene synthase
Procollagen type V 2Type VI collagen, α 3
Procollagen type XV
Ras-like GTP-binding protein Rem
Small proline-rich protein 2F
Snail homolog
Spermidine synthase
Squalene synthase
Tissue factor pathway inhibitor
Progesterone-responsiveEstrogen-responsive
Up-regulatedDown-regulated
Alkaline phosphataseArachidonate 15-lipoxygenaseBM-90/fibulin
AmphiregulinATFxEIG 180 (ethanol-induced gene)
Apg-2 (chaperone)Cytosolic adenylate kinaseGlutathione S-transferase, θ-2
Carbonic anhydrase IIGADD45 proteinHereditary hemochromatosis-like protein
Cathepsin FGlutamyl-tRNA synthetaseHoxd4
CCAAT/enhancer binding protein βGuanine nucleotide regulatory proteinHS1-binding protein 3
Chondroitin sulfate proteoglycan 2Heat shock protein, 105 kDaIntracisternal A particles
Claudin-7Hexokinase II, exon 1Leptin receptor
Complement C1q B chainIL-1 receptor, type 11Norrie disease homolog
Cyclin-dependent kinase inhibitor 1CMitochondrial stress-70 proteinP glycoprotein 3
Dickkopf-3NAD-dependent methylenetetrahydrofolate dimethyl cyclohydrolaseRas-related protein (DEXRAS1)
FollistatinThioether S-methyltransferase
Glutathione-S-transferaseNM23 metastatic-associated proteinXeroderma pigmentosum, complementation group C
Histidine decarboxylaseNuclear autoantigenic sperm protein
Hoxa 11p45 MAPK kinase
IGF binding protein-3Procollagen, type VI, α 2
IL-13 receptor, 2Protein kinase inhibitor p58
Keratin complex 1RAB geranylgeranyl transferase
LactotransferrinRAMP3
Leukocyte 12/15 lipoxygenaseRan GTPase
LRG-21RAN GTPase-activating protein 1
Membrane metalloendopeptidaseRNAse L inhibitor (Mu-RLI)
Metallothionein 1Small proline-rich protein 2F
Norrie disease homologSplicing factor, arginine/serine-rich 10
Osteoblast-specific factor 2Squalene epoxidase
Peptidylarginine deiminaseSqualene synthase
Procollagen type V 2Type VI collagen, α 3
Procollagen type XV
Ras-like GTP-binding protein Rem
Small proline-rich protein 2F
Snail homolog
Spermidine synthase
Squalene synthase
Tissue factor pathway inhibitor

The results of different mouse uterine microarray experiments were compared to identify commonly detected genes. Progesterone-responsive genes were pooled from several disparate approaches (253, 375, 377 ), whereas estrogen-induced genes were grouped by their response pattern (375, 379 ).

TABLE 1.

Identification of implantation-related genes in the mouse

Progesterone-responsiveEstrogen-responsive
Up-regulatedDown-regulated
Alkaline phosphataseArachidonate 15-lipoxygenaseBM-90/fibulin
AmphiregulinATFxEIG 180 (ethanol-induced gene)
Apg-2 (chaperone)Cytosolic adenylate kinaseGlutathione S-transferase, θ-2
Carbonic anhydrase IIGADD45 proteinHereditary hemochromatosis-like protein
Cathepsin FGlutamyl-tRNA synthetaseHoxd4
CCAAT/enhancer binding protein βGuanine nucleotide regulatory proteinHS1-binding protein 3
Chondroitin sulfate proteoglycan 2Heat shock protein, 105 kDaIntracisternal A particles
Claudin-7Hexokinase II, exon 1Leptin receptor
Complement C1q B chainIL-1 receptor, type 11Norrie disease homolog
Cyclin-dependent kinase inhibitor 1CMitochondrial stress-70 proteinP glycoprotein 3
Dickkopf-3NAD-dependent methylenetetrahydrofolate dimethyl cyclohydrolaseRas-related protein (DEXRAS1)
FollistatinThioether S-methyltransferase
Glutathione-S-transferaseNM23 metastatic-associated proteinXeroderma pigmentosum, complementation group C
Histidine decarboxylaseNuclear autoantigenic sperm protein
Hoxa 11p45 MAPK kinase
IGF binding protein-3Procollagen, type VI, α 2
IL-13 receptor, 2Protein kinase inhibitor p58
Keratin complex 1RAB geranylgeranyl transferase
LactotransferrinRAMP3
Leukocyte 12/15 lipoxygenaseRan GTPase
LRG-21RAN GTPase-activating protein 1
Membrane metalloendopeptidaseRNAse L inhibitor (Mu-RLI)
Metallothionein 1Small proline-rich protein 2F
Norrie disease homologSplicing factor, arginine/serine-rich 10
Osteoblast-specific factor 2Squalene epoxidase
Peptidylarginine deiminaseSqualene synthase
Procollagen type V 2Type VI collagen, α 3
Procollagen type XV
Ras-like GTP-binding protein Rem
Small proline-rich protein 2F
Snail homolog
Spermidine synthase
Squalene synthase
Tissue factor pathway inhibitor
Progesterone-responsiveEstrogen-responsive
Up-regulatedDown-regulated
Alkaline phosphataseArachidonate 15-lipoxygenaseBM-90/fibulin
AmphiregulinATFxEIG 180 (ethanol-induced gene)
Apg-2 (chaperone)Cytosolic adenylate kinaseGlutathione S-transferase, θ-2
Carbonic anhydrase IIGADD45 proteinHereditary hemochromatosis-like protein
Cathepsin FGlutamyl-tRNA synthetaseHoxd4
CCAAT/enhancer binding protein βGuanine nucleotide regulatory proteinHS1-binding protein 3
Chondroitin sulfate proteoglycan 2Heat shock protein, 105 kDaIntracisternal A particles
Claudin-7Hexokinase II, exon 1Leptin receptor
Complement C1q B chainIL-1 receptor, type 11Norrie disease homolog
Cyclin-dependent kinase inhibitor 1CMitochondrial stress-70 proteinP glycoprotein 3
Dickkopf-3NAD-dependent methylenetetrahydrofolate dimethyl cyclohydrolaseRas-related protein (DEXRAS1)
FollistatinThioether S-methyltransferase
Glutathione-S-transferaseNM23 metastatic-associated proteinXeroderma pigmentosum, complementation group C
Histidine decarboxylaseNuclear autoantigenic sperm protein
Hoxa 11p45 MAPK kinase
IGF binding protein-3Procollagen, type VI, α 2
IL-13 receptor, 2Protein kinase inhibitor p58
Keratin complex 1RAB geranylgeranyl transferase
LactotransferrinRAMP3
Leukocyte 12/15 lipoxygenaseRan GTPase
LRG-21RAN GTPase-activating protein 1
Membrane metalloendopeptidaseRNAse L inhibitor (Mu-RLI)
Metallothionein 1Small proline-rich protein 2F
Norrie disease homologSplicing factor, arginine/serine-rich 10
Osteoblast-specific factor 2Squalene epoxidase
Peptidylarginine deiminaseSqualene synthase
Procollagen type V 2Type VI collagen, α 3
Procollagen type XV
Ras-like GTP-binding protein Rem
Small proline-rich protein 2F
Snail homolog
Spermidine synthase
Squalene synthase
Tissue factor pathway inhibitor

The results of different mouse uterine microarray experiments were compared to identify commonly detected genes. Progesterone-responsive genes were pooled from several disparate approaches (253, 375, 377 ), whereas estrogen-induced genes were grouped by their response pattern (375, 379 ).

In humans, microarray technology has been used to examine uterine cell lines and explant cultures under conditions that simulate the periimplantation period (371, 382, 383). A more physiological application is based on estimates of the putative window of receptivity in women. Endometrial biopsies taken before and during this period have identified significant changes in a modest number of genes that may serve as markers for uterine receptivity in humans. Although some variation exists in their methods, two independent studies used the same microarray platforms for their analysis (384, 385). A comparison of the results shows differential regulation of a small number of genes around the time of uterine receptivity for implantation (Table 2). Interestingly, there are a number of these genes that are also differentially expressed in mouse models of implantation and uterine receptivity.

TABLE 2.

Identification of implantation-related genes in women

Up-regulated genesDown-regulated genes
Annexin IIAF1q
ApoDAutotaxin
Dickkopf-1 (hdkk-1)CAP2 adenylyl cyclase-associated protein
H2B/gCentromere protein-A (CENP-A)
hCPE-RChromosome 1-specific transcript
ID4 DNA-binding proteinCyclin B
IL-15Dipeptidyl aminopeptidase-like protein
IL-15 precursorEndothelin 3 (EDN3)
Igλ gene locusFibroblast activation protein
MammaglobinFrizzled-related protein frpHE
Metallothionein-IG (MTIG)GLI protein
NKG5 NK and T cell-specific genehFEN1
OsteopontinMatrilysin
Pyruvate carboxylaseMidline 1 fetal kidney isoform 3 (MID1)
TGFβ superfamily proteinMSX-2
Neuronal olfactomedin-related ER localized protein
RA-binding protein II (CRABP-II)
Serine/threonine kinase (STK-1)
TRAIL receptor 2
Up-regulated genesDown-regulated genes
Annexin IIAF1q
ApoDAutotaxin
Dickkopf-1 (hdkk-1)CAP2 adenylyl cyclase-associated protein
H2B/gCentromere protein-A (CENP-A)
hCPE-RChromosome 1-specific transcript
ID4 DNA-binding proteinCyclin B
IL-15Dipeptidyl aminopeptidase-like protein
IL-15 precursorEndothelin 3 (EDN3)
Igλ gene locusFibroblast activation protein
MammaglobinFrizzled-related protein frpHE
Metallothionein-IG (MTIG)GLI protein
NKG5 NK and T cell-specific genehFEN1
OsteopontinMatrilysin
Pyruvate carboxylaseMidline 1 fetal kidney isoform 3 (MID1)
TGFβ superfamily proteinMSX-2
Neuronal olfactomedin-related ER localized protein
RA-binding protein II (CRABP-II)
Serine/threonine kinase (STK-1)
TRAIL receptor 2

The results of two different human uterine microarray experiments were compared to identify commonly detected genes (384, 385 ). The objective of these two studies was to identify genes that are differentially expressed between prereceptive and receptive phases during the window of uterine receptivity for human implantation. Despite similarities in their approach, only a few genes were found to have common patterns of regulation during the putative window of human implantation. ApoD, Apolipoprotein D.

TABLE 2.

Identification of implantation-related genes in women

Up-regulated genesDown-regulated genes
Annexin IIAF1q
ApoDAutotaxin
Dickkopf-1 (hdkk-1)CAP2 adenylyl cyclase-associated protein
H2B/gCentromere protein-A (CENP-A)
hCPE-RChromosome 1-specific transcript
ID4 DNA-binding proteinCyclin B
IL-15Dipeptidyl aminopeptidase-like protein
IL-15 precursorEndothelin 3 (EDN3)
Igλ gene locusFibroblast activation protein
MammaglobinFrizzled-related protein frpHE
Metallothionein-IG (MTIG)GLI protein
NKG5 NK and T cell-specific genehFEN1
OsteopontinMatrilysin
Pyruvate carboxylaseMidline 1 fetal kidney isoform 3 (MID1)
TGFβ superfamily proteinMSX-2
Neuronal olfactomedin-related ER localized protein
RA-binding protein II (CRABP-II)
Serine/threonine kinase (STK-1)
TRAIL receptor 2
Up-regulated genesDown-regulated genes
Annexin IIAF1q
ApoDAutotaxin
Dickkopf-1 (hdkk-1)CAP2 adenylyl cyclase-associated protein
H2B/gCentromere protein-A (CENP-A)
hCPE-RChromosome 1-specific transcript
ID4 DNA-binding proteinCyclin B
IL-15Dipeptidyl aminopeptidase-like protein
IL-15 precursorEndothelin 3 (EDN3)
Igλ gene locusFibroblast activation protein
MammaglobinFrizzled-related protein frpHE
Metallothionein-IG (MTIG)GLI protein
NKG5 NK and T cell-specific genehFEN1
OsteopontinMatrilysin
Pyruvate carboxylaseMidline 1 fetal kidney isoform 3 (MID1)
TGFβ superfamily proteinMSX-2
Neuronal olfactomedin-related ER localized protein
RA-binding protein II (CRABP-II)
Serine/threonine kinase (STK-1)
TRAIL receptor 2

The results of two different human uterine microarray experiments were compared to identify commonly detected genes (384, 385 ). The objective of these two studies was to identify genes that are differentially expressed between prereceptive and receptive phases during the window of uterine receptivity for human implantation. Despite similarities in their approach, only a few genes were found to have common patterns of regulation during the putative window of human implantation. ApoD, Apolipoprotein D.

In addition to gene expression screening, numerous approaches have been taken to identify novel proteins that may be important for implantation. Overall uterine protein synthesis increases around the time of implantation and during artificially induced decidualization (386). The presence of unique proteins in implantation or decidual tissues was originally detected by immunological and radiolabeling experiments (387390). Two-dimensional gel electrophoresis has improved the ability to detect implantation-related proteins (391395). Currently, emerging techniques in proteomics have also led to applications in uterine biology (396398) and will likely resolve the presence and function of novel implantation-specific molecules.

VIII. Perspectives and Future Directions

Infertility and rapid population growth are two pressing global reproductive health issues. Events of preimplantation embryo development and uterine preparation for implantation are two major determinants of these concerns. Basic and clinical research to better understand these events will help alleviate problems of female infertility, improve fertility regulation in women, and lead to the development of new and improved contraceptive methods. Interactions between the heterogeneous cell types of the uterus and embryo in relation to endocrine, paracrine, juxtacrine, and autocrine factors during implantation are extremely complex (4). Thus, exploring and defining the molecular road map during the critical time of implantation necessitates well-thought out experimental designs in the context of both embryonic and uterine contributions to formulate a more meaningful blueprint. The objectives are not easily achievable in the human due to experimental difficulties, ethical considerations, and current restrictions on research with human embryos. Therefore, various animal models will continue to be used for studying embryo-uterine interactions during implantation. This may have relevance to human implantation. In addition, efforts should be directed to establish reliable in vitro systems, which are not currently available, to study implantation. However, experiments using endometrial biopsy samples to identify molecules associated with human uterine receptivity (window of implantation) during the menstrual cycle with changing estrogen and progesterone levels should continue to be pursued to obtain a better insight into human implantation.

Although a wealth of knowledge on the roles of growth factors, cytokines, homeotic genes, transcription factors, and lipid mediators in embryo-uterine interactions during implantation has been generated, their hierarchical blueprint in directing uterine and embryonic function during implantation remains to be deciphered. An arduous task before us is to unravel the intricate nature of the signaling pathways in implantation. We need to understand whether these pathways function independently, in parallel, or converge to a common signaling pathway to establish the network of cross-talk between the embryo and uterus that is necessary for implantation. Gene-targeting experiments in mice have identified a large number of genes that are important for female fertility; reproductive phenotypes of a select number of gene mutations are shown in Table 3. This list is not exhaustive and is likely to expand enormously with emergence of the rapidly evolving new technologies and results. Although some molecules involved in defining the pathways regulating implantation have been identified, our understanding of the implantation process is still far from complete. For example, many of the genes that are expressed in an implantation-specific manner and appear to be important for implantation cannot be studied in depth because deletion of these genes results in embryonic lethality. Uterine- or embryo-specific conditional knockout of genes of interest is urgently needed to better understand the definitive roles of these genes in uterine biology and implantation. Our failure to identify suitable uterine cell-specific promoters has been a hindrance to the achievement of this objective. One difficulty in identifying the critical roles of signaling molecules within a gene family is the redundant or compensatory functions of the gene products within the family.

TABLE 3.

Female reproductive performances in gene-targeted mouse models

GenePhenotypeRef.
Phenotypes related to embryo development
    α-CateninRetarded blastocyst development with disrupted trophectoderm leading to embryonic lethality401
    ATP-binding cassette transporter 1Defective steroid hormone synthesis; placental malformations and impaired embryo growth in mutant females leading to subfertility402
    Cannabinoid receptor (CB1)Asynchronous preimplantation development in female mutants leading to subfertility60
    E-cadherinFailure of blastocyst formation leading to early embryonic death150
    FGF-4Defective proliferation of inner cell mass causing embryonic demise403
    Heat shock transcription factor 1Normal ovulation and fertilization in female mutants, but failure of preimplantation embryo development (maternal effect) leading to female infertility404, 405
    Maternal effect gene (Mater)Defective preimplantation embryo development in female mutants leading to infertility406
    Nucleoplasmin 2Defective chromatin remodeling and preimplantation embryo development (subfertility)407
    Synaptonemal complex protein 3Defective meiotic chromosomal segmentation resulting in aneuploidy and embryonic death (subfertility)408
Phenotypes related to uterine biology and implantation
    ADAMTS-1Impaired fertilization with ovarian histological changes (uterine cysts) leading to subfertility409
    α-TocopherolFetal resorption due to vitamin D deficiency410
    BasiginDefective fertilization and implantation411, 412
    Centromere protein BDisrupted luminal and glandular uterine epithelium leading to subfertility (genetic background dependent)413
    Colony-stimulating factor-1Defective implantation and reduced fertility227, 414
    COX-2Impaired ovulation and fertilization; defective attachment reaction and decidualization associated with reduced angiogenic response186, 188
    Cytosolic phospholipase A2On-time implantation is deferred and gives rise to postimplantation defects leading to small litter size (maternal effect)54
    EGF-RImplantation or postimplantation failures of (−/−) embryos (genetic background dependent)222
    ERαOvarian cysts with uterine hypoplasia leading to infertility99, 415417
    Hmx3Failure of implantation (maternal effect)255
    HoxA 10Primarily defective decidualization leading to reduced fertility (maternal effect)248, 251, 418
    HoxA 11Lack of uterine glands with defective implantation and decidualization and infertility249, 250
    25-Hydroxyvitamin D 1α-hydroxylase enzymeUterine hypoplasia and absence of corpus luteum leading to infertility419
    IGF-IDefective ovarian and uterine functions with infertility420
    IL 11 receptor IL-11RαImpaired decidualization leading to infertility229
    LIFFailure in implantation and decidualization (maternal effect)228
    PPARδPlacental defects in the majority of (−/−) embryos leading to lethality; uterine defects causing subfertility190, 259
    PPARγPlacental defects leading to embryonic lethality263
    PRUnopposed estrogen action and uterine hyperplasia with failure in implantation and decidualization103
    Ubiquitin protein ligase E6APOvarian and uterine hypoplasia leading to subfertility421
    Wnt7aAbnormal oviduct and uterine development leading to infertility422
GenePhenotypeRef.
Phenotypes related to embryo development
    α-CateninRetarded blastocyst development with disrupted trophectoderm leading to embryonic lethality401
    ATP-binding cassette transporter 1Defective steroid hormone synthesis; placental malformations and impaired embryo growth in mutant females leading to subfertility402
    Cannabinoid receptor (CB1)Asynchronous preimplantation development in female mutants leading to subfertility60
    E-cadherinFailure of blastocyst formation leading to early embryonic death150
    FGF-4Defective proliferation of inner cell mass causing embryonic demise403
    Heat shock transcription factor 1Normal ovulation and fertilization in female mutants, but failure of preimplantation embryo development (maternal effect) leading to female infertility404, 405
    Maternal effect gene (Mater)Defective preimplantation embryo development in female mutants leading to infertility406
    Nucleoplasmin 2Defective chromatin remodeling and preimplantation embryo development (subfertility)407
    Synaptonemal complex protein 3Defective meiotic chromosomal segmentation resulting in aneuploidy and embryonic death (subfertility)408
Phenotypes related to uterine biology and implantation
    ADAMTS-1Impaired fertilization with ovarian histological changes (uterine cysts) leading to subfertility409
    α-TocopherolFetal resorption due to vitamin D deficiency410
    BasiginDefective fertilization and implantation411, 412
    Centromere protein BDisrupted luminal and glandular uterine epithelium leading to subfertility (genetic background dependent)413
    Colony-stimulating factor-1Defective implantation and reduced fertility227, 414
    COX-2Impaired ovulation and fertilization; defective attachment reaction and decidualization associated with reduced angiogenic response186, 188
    Cytosolic phospholipase A2On-time implantation is deferred and gives rise to postimplantation defects leading to small litter size (maternal effect)54
    EGF-RImplantation or postimplantation failures of (−/−) embryos (genetic background dependent)222
    ERαOvarian cysts with uterine hypoplasia leading to infertility99, 415417
    Hmx3Failure of implantation (maternal effect)255
    HoxA 10Primarily defective decidualization leading to reduced fertility (maternal effect)248, 251, 418
    HoxA 11Lack of uterine glands with defective implantation and decidualization and infertility249, 250
    25-Hydroxyvitamin D 1α-hydroxylase enzymeUterine hypoplasia and absence of corpus luteum leading to infertility419
    IGF-IDefective ovarian and uterine functions with infertility420
    IL 11 receptor IL-11RαImpaired decidualization leading to infertility229
    LIFFailure in implantation and decidualization (maternal effect)228
    PPARδPlacental defects in the majority of (−/−) embryos leading to lethality; uterine defects causing subfertility190, 259
    PPARγPlacental defects leading to embryonic lethality263
    PRUnopposed estrogen action and uterine hyperplasia with failure in implantation and decidualization103
    Ubiquitin protein ligase E6APOvarian and uterine hypoplasia leading to subfertility421
    Wnt7aAbnormal oviduct and uterine development leading to infertility422
TABLE 3.

Female reproductive performances in gene-targeted mouse models

GenePhenotypeRef.
Phenotypes related to embryo development
    α-CateninRetarded blastocyst development with disrupted trophectoderm leading to embryonic lethality401
    ATP-binding cassette transporter 1Defective steroid hormone synthesis; placental malformations and impaired embryo growth in mutant females leading to subfertility402
    Cannabinoid receptor (CB1)Asynchronous preimplantation development in female mutants leading to subfertility60
    E-cadherinFailure of blastocyst formation leading to early embryonic death150
    FGF-4Defective proliferation of inner cell mass causing embryonic demise403
    Heat shock transcription factor 1Normal ovulation and fertilization in female mutants, but failure of preimplantation embryo development (maternal effect) leading to female infertility404, 405
    Maternal effect gene (Mater)Defective preimplantation embryo development in female mutants leading to infertility406
    Nucleoplasmin 2Defective chromatin remodeling and preimplantation embryo development (subfertility)407
    Synaptonemal complex protein 3Defective meiotic chromosomal segmentation resulting in aneuploidy and embryonic death (subfertility)408
Phenotypes related to uterine biology and implantation
    ADAMTS-1Impaired fertilization with ovarian histological changes (uterine cysts) leading to subfertility409
    α-TocopherolFetal resorption due to vitamin D deficiency410
    BasiginDefective fertilization and implantation411, 412
    Centromere protein BDisrupted luminal and glandular uterine epithelium leading to subfertility (genetic background dependent)413
    Colony-stimulating factor-1Defective implantation and reduced fertility227, 414
    COX-2Impaired ovulation and fertilization; defective attachment reaction and decidualization associated with reduced angiogenic response186, 188
    Cytosolic phospholipase A2On-time implantation is deferred and gives rise to postimplantation defects leading to small litter size (maternal effect)54
    EGF-RImplantation or postimplantation failures of (−/−) embryos (genetic background dependent)222
    ERαOvarian cysts with uterine hypoplasia leading to infertility99, 415417
    Hmx3Failure of implantation (maternal effect)255
    HoxA 10Primarily defective decidualization leading to reduced fertility (maternal effect)248, 251, 418
    HoxA 11Lack of uterine glands with defective implantation and decidualization and infertility249, 250
    25-Hydroxyvitamin D 1α-hydroxylase enzymeUterine hypoplasia and absence of corpus luteum leading to infertility419
    IGF-IDefective ovarian and uterine functions with infertility420
    IL 11 receptor IL-11RαImpaired decidualization leading to infertility229
    LIFFailure in implantation and decidualization (maternal effect)228
    PPARδPlacental defects in the majority of (−/−) embryos leading to lethality; uterine defects causing subfertility190, 259
    PPARγPlacental defects leading to embryonic lethality263
    PRUnopposed estrogen action and uterine hyperplasia with failure in implantation and decidualization103
    Ubiquitin protein ligase E6APOvarian and uterine hypoplasia leading to subfertility421
    Wnt7aAbnormal oviduct and uterine development leading to infertility422
GenePhenotypeRef.
Phenotypes related to embryo development
    α-CateninRetarded blastocyst development with disrupted trophectoderm leading to embryonic lethality401
    ATP-binding cassette transporter 1Defective steroid hormone synthesis; placental malformations and impaired embryo growth in mutant females leading to subfertility402
    Cannabinoid receptor (CB1)Asynchronous preimplantation development in female mutants leading to subfertility60
    E-cadherinFailure of blastocyst formation leading to early embryonic death150
    FGF-4Defective proliferation of inner cell mass causing embryonic demise403
    Heat shock transcription factor 1Normal ovulation and fertilization in female mutants, but failure of preimplantation embryo development (maternal effect) leading to female infertility404, 405
    Maternal effect gene (Mater)Defective preimplantation embryo development in female mutants leading to infertility406
    Nucleoplasmin 2Defective chromatin remodeling and preimplantation embryo development (subfertility)407
    Synaptonemal complex protein 3Defective meiotic chromosomal segmentation resulting in aneuploidy and embryonic death (subfertility)408
Phenotypes related to uterine biology and implantation
    ADAMTS-1Impaired fertilization with ovarian histological changes (uterine cysts) leading to subfertility409
    α-TocopherolFetal resorption due to vitamin D deficiency410
    BasiginDefective fertilization and implantation411, 412
    Centromere protein BDisrupted luminal and glandular uterine epithelium leading to subfertility (genetic background dependent)413
    Colony-stimulating factor-1Defective implantation and reduced fertility227, 414
    COX-2Impaired ovulation and fertilization; defective attachment reaction and decidualization associated with reduced angiogenic response186, 188
    Cytosolic phospholipase A2On-time implantation is deferred and gives rise to postimplantation defects leading to small litter size (maternal effect)54
    EGF-RImplantation or postimplantation failures of (−/−) embryos (genetic background dependent)222
    ERαOvarian cysts with uterine hypoplasia leading to infertility99, 415417
    Hmx3Failure of implantation (maternal effect)255
    HoxA 10Primarily defective decidualization leading to reduced fertility (maternal effect)248, 251, 418
    HoxA 11Lack of uterine glands with defective implantation and decidualization and infertility249, 250
    25-Hydroxyvitamin D 1α-hydroxylase enzymeUterine hypoplasia and absence of corpus luteum leading to infertility419
    IGF-IDefective ovarian and uterine functions with infertility420
    IL 11 receptor IL-11RαImpaired decidualization leading to infertility229
    LIFFailure in implantation and decidualization (maternal effect)228
    PPARδPlacental defects in the majority of (−/−) embryos leading to lethality; uterine defects causing subfertility190, 259
    PPARγPlacental defects leading to embryonic lethality263
    PRUnopposed estrogen action and uterine hyperplasia with failure in implantation and decidualization103
    Ubiquitin protein ligase E6APOvarian and uterine hypoplasia leading to subfertility421
    Wnt7aAbnormal oviduct and uterine development leading to infertility422

Strategies comparing global gene expression profiles between the implantation and interimplantation sites have identified novel genes in the implantation process. Thus, a genome-wide screening approach coupled with functional assays will help elucidate these complex signaling pathways. In addition, further experiments should be pursued to compare gene expression patterns between the uterine-receptive and nonreceptive phases, and between the active and dormant blastocysts under defined experimental conditions. The results obtained from these experiments may help uncover new signaling molecules and pathways not previously identified. The application of proteomics is also likely to provide information regarding interactions among various molecular pathways in specifying the molecular road map to implantation. Another area of research that deserves particular attention is the identification of embryonic signaling molecules that influence uterine functions for implantation. Although chorionic gonadotropins in primates are well known for their role in pregnancy establishment, it is not yet clear whether they also function as implantation initiators. The situation is different in large animals in which embryo-derived interferons and estrogens are known to function as important signaling molecules for pregnancy recognition (12, 399, 400). These studies have been possible due to the availability of relatively large amounts of blastocyst tissue in these species. In other species, including rodents and humans, the most limiting factor has been the availability of an adequate amount of tissue for biochemical and molecular biology experiments. With the advent of microscale proteomics and genomic approaches, it is hoped that more information on embryonic signals is likely to be forthcoming.

Finally, although the mechanics and cellular architecture of the implantation process vary, certain basic similarities do exist among species. For example, implantation occurs at the blastocyst stage, there is a defined window of uterine receptivity for implantation, a reciprocal interaction between the blastocyst and the uterus is essential for implantation, and a localized increase in uterine vascular permeability occurs at the site of the blastocyst during the attachment reaction. Thus, recognition and characterization of signaling pathways in these steps may give rise to a unifying scheme relevant to understanding the mechanism of human implantation.

Acknowledgement

We thank Ben Danzo for his critical reading and editing of the manuscript and Stan Fernald of the Cell Imaging Facility at the Center for Reproductive Sciences at the University of Kansas Medical Center for the artwork. We apologize for unintended omission of any relevant references.

Studies by the authors incorporated in this review were supported by National Institutes of Health Grants HD29968, HD12304, HD33994, DA 06668 (to S.K.De.), HD 40810 (to H.L.), HD37830 (to S.K.Da.), HD40221 (to J.R.), and HD37394 (to B.C.P.). S.K.De. is recipient of MERIT Awards from the National Institute of Child Health and Human Development and the National Institute on Drug Abuse. H.W. is Lalor Foundation postdoctoral fellow.

Abbreviations

     
  • 2-AG,

    2-Arachidonoylglycerol;

  •  
  • Ang1,

    angiopoietin-1;

  •  
  • ARNT,

    aryl hydrocarbon nuclear translocator;

  •  
  • BMP,

    bone morphogenetic protein;

  •  
  • CBP,

    cAMP response element-binding protein (CREB)-binding protein;

  •  
  • cdk,

    cyclin-dependent kinase;

  •  
  • CKI,

    cyclin-dependent kinase inhibitor;

  •  
  • COX-2,

    cyclooxygenase-2;

  •  
  • DDRT-PCR,

    differential display RT-PCR;

  •  
  • ECM,

    extracellular matrix;

  •  
  • EGF,

    epidermal growth factor;

  •  
  • EGF-R,

    EGF receptor;

  •  
  • ER,

    estrogen receptor;

  •  
  • ERE,

    estrogen response element;

  •  
  • FGF,

    fibroblast growth factor;

  •  
  • FLK1,

    fetal liver kinase 1;

  •  
  • FLT1,

    fms-like tyrosine kinase 1;

  •  
  • H2,

    histamine type 2 receptor;

  •  
  • HB-EGF,

    heparin-binding EFG-like growth factor;

  •  
  • HDC,

    histidine decarboxylase;

  •  
  • HIF,

    hypoxia-inducible factor;

  •  
  • ICM,

    inner cell mass;

  •  
  • IHH,

    Indian hedgehog;

  •  
  • LIF,

    leukemia inhibitory factor;

  •  
  • MMP,

    matix metalloproteinase;

  •  
  • NRP1,

    Neuropilin-1;

  •  
  • 4-OH-E2,

    4-hydroxy-17β-estradiol;

  •  
  • PBP,

    PPAR-binding protein;

  •  
  • PDZ,

    primary decidual zone;

  •  
  • PG,

    prostaglandin;

  •  
  • PGI2,

    prostacyclin;

  •  
  • PlGF,

    placental growth factor;

  •  
  • PPAR,

    peroxisome proliferator-activated receptor;

  •  
  • PR,

    progesterone receptor;

  •  
  • PRIP,

    PPAR-interacting protein;

  •  
  • PTC,

    Patched;

  •  
  • RGD,

    Arg-Gly-Asp sequence;

  •  
  • RIP,

    receptor interacting protein;

  •  
  • RXR,

    retinoid X receptor;

  •  
  • SAGE,

    serial analysis of gene expression;

  •  
  • SDZ,

    secondary decidual zone;

  •  
  • SRC,

    steroid receptor coactivator;

  •  
  • TIF,

    transcriptional intermediary factor;

  •  
  • TIMP,

    tissue inhibitor of MMPs;

  •  
  • TR,

    thyroid hormone receptor;

  •  
  • VEGF,

    vascular endothelial growth factor.

1

Gardner
RL
,
Papaioannou
VE
1975
Differentiation in the trophectoderm and inner cell mass. In:
Balls
M
,
Wild
AE
, eds. The early development of mammals.
London
:
Cambridge University Press
;
107
132

2

Renfree
MB
1982
Implantation and placentation. In:
Austin
CR
,
Short
RV
, eds. Reproduction in mammals.
Cambridge, UK
:
Cambridge University Press
;
26
69

3

Carson
DD
,
Bagchi
I
,
Dey
SK
,
Enders
AC
,
Fazleabas
AT
,
Lessey
BA
,
Yoshinaga
K
2000
Embryo implantation.
Dev Biol
223
:
217
237

4

Paria
BC
,
Reese
J
,
Das
SK
,
Dey
SK
2002
Deciphering the cross-talk of implantation: advances and challenges.
Science
296
:
2185
2188

5

Lim
H
,
Song
H
,
Paria
BC
,
Reese
J
,
Das
SK
,
Dey
SK
2002
Molecules in blastocyst implantation: uterine and embryonic perspectives.
Vitam Horm
64
:
43
76

6

Norwitz
ER
,
Schust
DJ
,
Fisher
SJ
2001
Implantation and the survival of early pregnancy.
N Engl J Med
345
:
1400
1408

7

Giudice
LC
1999
Potential biochemical markers of uterine receptivity.
Hum Reprod
14(Suppl 2)
:
3
16

8

Artley
JK
,
Braude
PR
,
Johnson
MH
1992
Gene activity and cleavage arrest in human pre-embryos.
Hum Reprod
7
:
1014
1021

9

Hogan
B
,
Constantini
F
,
Lacy
E
1994
Manipulating the mouse embryo: a laboratory manual, 2nd ed.
Cold Spring Harbor, NY
:
Cold Spring Harbor Press

10

McLaren
A
1990
The embryo. In: Austin CR, Short RV eds.
Reproduction in mammals. Cambridge, UK
:
Cambridge University Press; 1–26

11

Petraglia
F
,
Santuz
M
,
Florio
P
,
Simoncini
T
,
Luisi
S
,
Plaino
L
,
Genazzani
AR
,
Genazzani
AD
,
Volpe
A
1998
Paracrine regulation of human placenta: control of hormonogenesis.
J Reprod Immunol
39
:
221
233

12

Roberts
RM
,
Ealy
AD
,
Alexenko
AP
,
Han
CS
,
Ezashi
T
1999
Trophoblast interferons.
Placenta
20
:
259
264

13

Sherman
MI
1978
Implantation of mouse blastocysts in vitro. In:
Daniels
JC
, ed. Methods in mammalian reproduction.
New York
:
Academic Press
;
247
257

14

McLaren
A
1971
Blastocysts in the mouse uterus: the effect of ovariectomy, progesterone and oestrogen.
J Endocrinol
50
:
515
526

15

Bowman
P
,
McLaren
A
1970
Cleavage rate of mouse embryos in vivo and in vitro.
J Embryol Exp Morphol
24
:
203
207

16

Roblero
LS
,
Garavagno
AC
1979
Effect of oestradiol-17 β and progesterone on oviductal transport and early development of mouse embryos.
J Reprod Fertil
57
:
91
95

17

Paria
BC
,
Lim
H
,
Wang
XN
,
Liehr
J
,
Das
SK
,
Dey
SK
1998
Coordination of differential effects of primary estrogen and catecholestrogen on two distinct targets mediates embryo implantation in the mouse.
Endocrinology
139
:
5235
5246

18

Paria
BC
,
Dey
SK
1990
Preimplantation embryo development in vitro: cooperative interactions among embryos and role of growth factors.
Proc Natl Acad Sci USA
87
:
4756
4760

19

Paria
BC
,
Lim
H
,
Das
SK
,
Reese
J
,
Dey
SK
2000
Molecular signaling in uterine receptivity for implantation.
Semin Cell Dev Biol
11
:
67
76

20

Stewart
CL
,
Cullinan
EB
1997
Preimplantation development of the mammalian embryo and its regulation by growth factors.
Dev Genet
21
:
91
101

21

Enders
AC
,
Schlafke
S
1967
A morphological analysis of the early implantation stages in the rat.
Am J Anat
120
:
185
226

22

Enders
AC
,
Schlafke
S
1969
Cytological aspects of trophoblast-uterine interaction in early implantation.
Am J Anat
125
:
1
29

23

Bonnet
R
1884
Beitrage zur embryologie der wiederkauer, gewonnen am schafei.
Arch Anat Physiol
8
:
170
230

24

Schlafke
S
,
Enders
AC
1975
Cellular basis of interaction between trophoblast and uterus at implantation.
Biol Reprod
12
:
41
65

25

Psychoyos
A
1973
Endocrine control of egg implantation. In:
Greep
RO
,
Astwood
EG
,
Geiger
SR
, eds. Handbook of physiology.
Washington, DC
:
American Physiology Society
;
187
215

26

Das
SK
,
Wang
XN
,
Paria
BC
,
Damm
D
,
Abraham
JA
,
Klagsbrun
M
,
Andrews
GK
,
Dey
SK
1994
Heparin-binding EGF-like growth factor gene is induced in the mouse uterus temporally by the blastocyst solely at the site of its apposition: a possible ligand for interaction with blastocyst EGF-receptor in implantation.
Development
120
:
1071
1083

27

Hoos
PC
,
Hoffman
LH
1980
Temporal aspects of rabbit uterine vascular and decidual responses to blastocyst stimulation.
Biol Reprod
23
:
453
459

28

Enders
AC
,
Schlafke
S
1986
Implantation in nonhuman primates and in the human.
Comp Primate Biol
3
:
453
459

29

Enders
AC
,
Lopata
A
1999
Implantation in the marmoset monkey: expansion of the early implantation site.
Anat Rec
256
:
279
299

30

Dey
SK
1996
Implantation. In:
Adashi
EY
,
Rock
JA
,
Rosenwaks
Z
, eds. Reproductive endocrinology, surgery and technology.
New York
:
Lippincott-Raven
;
421
434

31

Finn
CA
,
Porter
DG
1975
The uterus.
Acton, MA
:
Publishing Sciences Group, Inc
.

32

Kirby
DRS
1971
Blastocyst-uterine relationship before and during implantation. In:
Blandau
RJ
, ed. The biology of blastocysts.
Chicago
:
The University of Chicago Press
;
393
410

33

Mead
RA
1993
Embryonic diapause in vertebrates.
J Exp Zool
266
:
629
641

34

Yoshinaga
K
,
Adams
CE
1966
Delayed implantation in the spayed, progesterone treated adult mouse.
J Reprod Fertil
12
:
593
595

35

Yoshinaga
K
,
Hosi
T
1958
On the delayed implantation in lactating pregnant rat. I. The effect of estrogen.
Jpn J Anim Reprod
3
:
93
94

36

McLaren
A
1968
A study of blastocysts during delay and subsequent implantation in lactating mice.
J Endocrinol
42
:
453
463

37

Aitken
RJ
1977
Embryonic diapause. In:
Johnson
MH
, ed. Development in mammals.
Amsterdam
:
North Holland Publishing Co
;
307
359

38

Daniel Jr
JC
1973
Dormant embryos of mammals.
Bio Sci
20
:
411
415

39

McLaren
A
1973
Blastocyst activation. In:
Segal
SJ
,
Crozier
R
,
Corfman
PA
,
Condliffe
PG
, eds. The regulation of mammalian reproduction.
Springfield, IL
:
Charles C. Thomas
;
321
328

40

Mead
RA
1981
Delayed implantation in mustelids, with special emphasis on the spotted skunk.
J Reprod Fertil Suppl
29
:
11
24

41

Psychoyos
A
1973
Hormonal control of ovoimplantation.
Vitam Horm
31
:
201
256

42

Yoshinaga
K
1988
Uterine receptivity for blastocyst implantation.
Ann NY Acad Sci
541
:
424
431

43

Paria
BC
,
Huet
H
,
Dey
SK
1993
Blastocyst’s state of activity determines the “window” of implantation in the receptive mouse uterus.
Proc Natl Acad Sci USA
90
:
10159
10162

44

Heap
RB
,
Deanesly
R
1967
The increase in plasma progesterone levels in the pregnant guinea-pig and its possible significance.
J Reprod Fertil
14
:
339
341

45

Harper
MJ
,
Dowd
D
,
Elliott
AS
1969
Implantation and embryonic development in the ovariectomized-adrenalectomized hamster.
Biol Reprod
1
:
253
257

46

Kwun
JK
,
Emmens
CW
1974
Hormonal requirements for implantation and pregnancy in the ovariectomized rabbit.
Aust J Biol Sci
27
:
275
283

47

McCormack
JT
,
Greenwald
GS
1974
Evidence for a preimplantation rise in oestradiol-17β levels on day 4 of pregnancy in the mouse.
J Reprod Fertil
41
:
297
301

48

Heap
RB
,
Flint
AP
,
Hartmann
PE
,
Gadsby
JE
,
Staples
LD
,
Ackland
N
,
Hamon
M
1981
Oestrogen production in early pregnancy.
J Endocrinol
89 (Suppl)
:
77P
94P

49

Stromstedt
M
,
Keeney
DS
,
Waterman
MR
,
Paria
BC
,
Conley
AJ
,
Dey
SK
1996
Preimplantation mouse blastocysts fail to express CYP genes required for estrogen biosynthesis.
Mol Reprod Dev
43
:
428
436

50

Huet
H
,
Andrews
GK
,
Dey
SK
1989
Cell type-specific localization of c-myc protein in the mouse uterus: modulation by steroid hormones and analysis of the periimplantation period.
Endocrinology
125
:
1683
1690

51

Lundkvist
O
,
Nilsson
BO
1982
Endometrial ultrastructure in the early uterine response to blastocysts and artificial deciduogenic stimuli in rats.
Cell Tissue Res
225
:
355
364

52

Paria
BC
,
Ma
W
,
Tan
J
,
Raja
S
,
Das
SK
,
Dey
SK
,
Hogan
BL
2001
Cellular and molecular responses of the uterus to embryo implantation can be elicited by locally applied growth factors.
Proc Natl Acad Sci USA
98
:
1047
1052

53

Huet
H
,
Dey
SK
1990
Requirement for progesterone priming and its long-term effects on implantation in the mouse.
Proc Soc Exp Biol Med
193
:
259
263

54

Song
H
,
Lim
H
,
Paria
BC
,
Matsumoto
H
,
Swift
LL
,
Morrow
J
,
Bonventre
JV
,
Dey
SK
2002
Cytosolic phospholipase A2α is crucial [correction of A2α deficiency is crucial] for ’on-time’ embryo implantation that directs subsequent development.
Development [Erratum (2002)
129
:
3761
]129:2879–2889

55

Ma
WG
,
Song
H
,
Das
SK
,
Paria
BC
,
Dey
SK
2003
Estrogen is a critical determinant that specifies the duration of the window of uterine receptivity for implantation.
Proc Natl Acad Sci USA
100
:
2963
2968

56

Paria
BC
,
Das
SK
,
Andrews
GK
,
Dey
SK
1993
Expression of the epidermal growth factor receptor gene is regulated in mouse blastocysts during delayed implantation.
Proc Natl Acad Sci USA
90
:
55
59

57

Raab
G
,
Kover
K
,
Paria
BC
,
Dey
SK
,
Ezzell
RM
,
Klagsbrun
M
1996
Mouse preimplantation blastocysts adhere to cells expressing the transmembrane form of heparin-binding EGF-like growth factor.
Development
122
:
637
645

58

Lim
H
,
Das
SK
,
Dey
SK
1998
erbB genes in the mouse uterus: cell-specific signaling by epidermal growth factor (EGF) family of growth factors during implantation.
Dev Biol
204
:
97
110

59

Zhao
X
,
Ma
W
,
Das
SK
,
Dey
SK
,
Paria
BC
2000
Blastocyst H(2) receptor is the target for uterine histamine in implantation in the mouse.
Development
127
:
2643
2651

60

Paria
BC
,
Song
H
,
Wang
X
,
Schmid
PC
,
Krebsbach
RJ
,
Schmid
HH
,
Bonner
TI
,
Zimmer
A
,
Dey
SK
2001
Dysregulated cannabinoid signaling disrupts uterine receptivity for embryo implantation.
J Biol Chem
276
:
20523
20528

61

Tsai
MJ
,
O’Malley
BW
1994
Molecular mechanisms of action of steroid/thyroid receptor superfamily members.
Annu Rev Biochem
63
:
451
486

62

Beato
M
,
Herrlich
P
,
Schutz
G
1995
Steroid hormone receptors: many actors in search of a plot.
Cell
83
:
851
857

63

Kuiper
GG
,
Enmark
E
,
Pelto
H
,
Nilsson
S
,
Gustafsson
JA
1996
Cloning of a novel receptor expressed in rat prostate and ovary.
Proc Natl Acad Sci USA
93
:
5925
5930

64

Couse
JF
,
Korach
KS
1999
Estrogen receptor null mice: what have we learned and where will they lead us?
Endocr Rev
20
:
358
417

65

Stancel
GM
,
Boettger
T
,
Chiappetta
C
,
Hyder
SM
,
Kirkland
JL
,
Murthy
L
,
Loose
M
1995
Toxicity of endogenous and environmental estrogens: what is the role of elemental interactions?
Environ Health Perspect
103(Suppl 7)
:
29
33

66

Klein
H
,
Ryffel
GU
,
Heitlinger
E
,
Cato
AC
1988
A 13 bp palindrome is a functional estrogen responsive element and interacts specifically with estrogen receptor.
Nucleic Acids Res
16
:
647
663

67

Aumais
JP
,
Lee
HS
,
DeGannes
C
,
Horsford
J
,
White
JH
1996
Function of directly repeated half-sites as response elements for steroid hormone receptors.
J Biol Chem
271
:
12568
12577

68

Umayahara
Y
,
Kawamori
R
,
Watada
H
,
Imano
E
,
Iwama
N
,
Morishima
T
,
Yamasaki
Y
,
Kajimoto
Y
,
Kamada
T
1994
Estrogen regulation of the insulin-like growth factor I gene transcription involves an AP-1 enhancer.
J Biol Chem
269
:
16433
16442

69

Sukovich
DA
,
Mukherjee
R
,
Benfield
PA
1994
A novel, cell-type-specific mechanism for estrogen receptor-mediated gene activation in the absence of an estrogen-responsive element.
Mol Cell Biol
14
:
7134
7143

70

Krishnan
V
,
Wang
X
,
Safe
S
1994
Estrogen receptor-Sp1 complexes mediate estrogen-induced cathepsin D gene expression in MCF-7 human breast cancer cells.
J Biol Chem
269
:
15912
15917

71

Kato
S
,
Endoh
H
,
Masuhiro
Y
,
Kitamoto
T
,
Uchiyama
S
,
Sasaki
H
,
Masushige
S
,
Gotoh
Y
,
Nishida
E
,
Kawashima
H
1995
Activation of the estrogen receptor through phosphorylation by mitogen-activated protein kinase.
Science
270
:
1491
1494

72

McKenna
NJ
,
O’Malley
BW
2002
Combinatorial control of gene expression by nuclear receptors and coregulators.
Cell
108
:
465
474

73

McKenna
NJ
,
O’Malley
BW
2000
An issue of tissues: divining the split personalities of selective estrogen receptor modulators.
Nat Med
6
:
960
962

74

Stewart
PJ
,
Webster
RA
1983
Intrauterine injection of cholera toxin induces estrogen-like uterine growth.
Biol Reprod
29
:
671
679

75

Aronica
SM
,
Kraus
WL
,
Katzenellenbogen
BS
1994
Estrogen action via the cAMP signaling pathway: stimulation of adenylate cyclase and cAMP-regulated gene transcription.
Proc Natl Acad Sci USA
91
:
8517
8521

76

Rajkumar
K
1993
Effect of protein kinase-C inhibitor on estradiol-induced deoxyribonucleic acid synthesis in rats.
Steroids
58
:
100
105

77

Pietras
RJ
,
Szego
CM
1977
Specific binding sites for oestrogen at the outer surfaces of isolated endometrial cells.
Nature
265
:
69
72

78

Razandi
M
,
Pedram
A
,
Park
ST
,
Levin
ER
2003
Proximal events in signaling by plasma membrane estrogen receptors.
J Biol Chem
278
:
2701
2712

79

Jobling
S
,
Reynolds
T
,
White
R
,
Parker
MG
,
Sumpter
JP
1995
A variety of environmentally persistent chemicals, including some phthalate plasticizers, are weakly estrogenic.
Environ Health Perspect
103
:
582
587

80

Whitten
PL
,
Naftolin
F
1990
Dietary estrogens: a biologically active background for estrogen action. In:
Hochberg
RB
,
Naftolin
F
, eds. The new biology of steroid hormones.
New York
:
Raven Press
;
155
167

81

Kuiper
GG
,
Carlsson
B
,
Grandien
K
,
Enmark
E
,
Haggblad
J
,
Nilsson
S
,
Gustafsson
JA
1997
Comparison of the ligand binding specificity and transcript tissue distribution of estrogen receptors α and β.
Endocrinology
138
:
863
870

82

Cowley
SM
,
Hoare
S
,
Mosselman
S
,
Parker
MG
1997
Estrogen receptors α and β form heterodimers on DNA.
J Biol Chem
272
:
19858
19862

83

Nelson
JA
,
Struck
RF
,
James
R
1978
Estrogenic activities of chlorinated hydrocarbons.
J Toxicol Environ Health
4
:
325
339

84

Robinson
AK
,
Mukku
VR
,
Stancel
GM
1985
Analysis and characterization of estrogenic xenobiotics and natural products. In:
McLachlan
JA
, ed. Estrogens in the environment.
New York
:
Elsevier Science Publishing, Inc.
;
107
115

85

Sirois
J
,
Simmons
DL
,
Richards
JS
1992
Hormonal regulation of messenger ribonucleic acid encoding a novel isoform of prostaglandin endoperoxide H synthase in rat preovulatory follicles. Induction in vivo and in vitro.
J Biol Chem
267
:
11586
11592

86

Kupfer
D
,
Bulger
WH
1990
Inactivation of the uterine estrogen receptor binding of estradiol during P-450 catalyzed metabolism of chlorotrianisene (TACE). Speculation that TACE antiestrogenic activity involves covalent binding to the estrogen receptor.
FEBS Lett
261
:
59
62

87

Meyers
CY
,
Kolb
VM
,
Gass
GH
,
Rao
BR
,
Roos
CF
,
Dandliker
WB
1988
Doisynolic-type acids—uterotropically potent estrogens which compete poorly with estradiol for cytosolic estradiol receptors.
J Steroid Biochem
31
:
393
404

88

Meyers
CY
,
Lutfi
HG
,
Adler
S
1997
Transcriptional regulation of estrogen-responsive genes by non-steroidal estrogens: doisynolic and allenolic acids.
J Steroid Biochem Mol Biol
62
:
477
489

89

Johnson
DC
,
Sen
M
,
Kogo
H
,
Dey
SK
1990
Initiation of embryo implantation and maintenance of early pregnancy in the rat by chlordecone (Kepone).
Proc Soc Exp Biol Med
195
:
44
50

90

Cummings
AM
1997
Methoxychlor as a model for environmental estrogens.
Crit Rev Toxicol
27
:
367
379

91

Das
SK
,
Taylor
JA
,
Korach
KS
,
Paria
BC
,
Dey
SK
,
Lubahn
DB
1997
Estrogenic responses in estrogen receptor-α deficient mice reveal a distinct estrogen signaling pathway.
Proc Natl Acad Sci USA
94
:
12786
12791

92

Das
SK
,
Tan
J
,
Raja
S
,
Halder
J
,
Paria
BC
,
Dey
SK
2000
Estrogen targets genes involved in protein processing, calcium homeostasis, and Wnt signaling in the mouse uterus independent of estrogen receptor-α and -β.
J Biol Chem
275
:
28834
28842

93

Das
SK
,
Tan
J
,
Johnson
DC
,
Dey
SK
1998
Differential spatiotemporal regulation of lactoferrin and progesterone receptor genes in the mouse uterus by primary estrogen, catechol estrogen, and xenoestrogen.
Endocrinology
139
:
2905
2915

94

Das
SK
,
Paria
BC
,
Johnson
DC
,
Dey
SK
1997
Embryo-uterine interactions during implantation: potential sites of interference by environmental toxins. In:
Boekelheide
K
,
Chapin
RE
,
Hoyer
PB
,
Harris
C
, eds. Comprehensive toxicology.
New York
:
Elsevier Science Ltd
;
317
328

95

Ghosh
D
,
Taylor
JA
,
Green
JA
,
Lubahn
DB
1999
Methoxychlor stimulates estrogen-responsive messenger ribonucleic acids in mouse uterus through a non-estrogen receptor (non-ER) α and non-ER β mechanism.
Endocrinology
140
:
3526
3533

96

Couse
JF
,
Curtis
SW
,
Washburn
TF
,
Lindzey
J
,
Golding
TS
,
Lubahn
DB
,
Smithies
O
,
Korach
KS
1995
Analysis of transcription and estrogen insensitivity in the female mouse after targeted disruption of the estrogen receptor gene.
Mol Endocrinol
9
:
1441
1454

97

Chen
D
,
Ganapathy
P
,
Zhu
LJ
,
Xu
X
,
Li
Q
,
Bagchi
IC
,
Bagchi
MK
1999
Potential regulation of membrane trafficking by estrogen receptor α via induction of rab11 in uterine glands during implantation.
Mol Endocrinol
13
:
993
1004

98

Tan
J
,
Paria
BC
,
Dey
SK
,
Das
SK
1999
Differential uterine expression of estrogen and progesterone receptors correlates with uterine preparation for implantation and decidualization in the mouse.
Endocrinology
140
:
5310
5321

99

Lubahn
DB
,
Moyer
JS
,
Golding
TS
,
Couse
JF
,
Korach
KS
,
Smithies
O
1993
Alteration of reproductive function but not prenatal sexual development after insertional disruption of the mouse estrogen receptor gene.
Proc Natl Acad Sci USA
90
:
11162
11166

100

Curtis
H
,
Goulding
EH
,
Eddy
EM
,
Korach
KS
2002
Studies using the estrogen receptor α knockout uterus demonstrate that implantation but not decidualization-associated signaling is estrogen dependent.
Biol Reprod
67
:
1268
1277

101

Paria
BC
,
Tan
J
,
Lubahn
DB
,
Dey
SK
,
Das
SK
1999
Uterine decidual response occurs in estrogen receptor-α-deficient mice.
Endocrinology
140
:
2704
2710

102

Curtis
SW
,
Clark
J
,
Myers
P
,
Korach
KS
1999
Disruption of estrogen signaling does not prevent progesterone action in the estrogen receptor α knockout mouse uterus.
Proc Natl Acad Sci USA
96
:
3646
3651

103

Lydon
JP
,
DeMayo
FJ
,
Funk
CR
,
Mani
SK
,
Hughes
AR
,
Montgomery
CA
,
Shyamala
G
,
Conneely
OM
,
O’Malley
BW
1995
Mice lacking progesterone receptor exhibit pleiotropic reproductive abnormalities.
Genes Dev
9
:
2266
2278

104

Mulac
J
,
Mullinax
RA
,
DeMayo
FJ
,
Lydon
JP
,
Conneely
OM
2000
Subgroup of reproductive functions of progesterone mediated by progesterone receptor-B isoform.
Science
289
:
1751
1754

105

Hou
Q
,
Paria
BC
,
Mui
C
,
Dey
SK
,
Gorski
J
1996
Immunolocalization of estrogen receptor protein in the mouse blastocyst during normal and delayed implantation.
Proc Natl Acad Sci USA
93
:
2376
2381

106

Kowalski
AA
,
Graddy
LG
,
Vale
C
,
Choi
I
,
Katzenellenbogen
BS
,
Simmen
FA
,
Simmen
RC
2002
Molecular cloning of porcine estrogen receptor-β complementary DNAs and developmental expression in periimplantation embryos.
Biol Reprod
66
:
760
769

107

Hiroi
H
,
Momoeda
M
,
Inoue
S
,
Tsuchiya
F
,
Matsumi
H
,
Tsutsumi
O
,
Muramatsu
M
,
Taketani
Y
1999
Stage-specific expression of estrogen receptor subtypes and estrogen responsive finger protein in preimplantational mouse embryos.
Endocr J
46
:
153
158

108

Shen
Z
,
Liu
J
,
Wells
RL
,
Elkind
MM
1994
cDNA cloning, sequence analysis, and induction by aryl hydrocarbons of a murine cytochrome P450 gene, Cyp1b1.
DNA Cell Biol
13
:
763
769

109

Aplin
JD
1997
Adhesion molecules in implantation.
Rev Reprod
2
:
84
93

110

Kimber
SJ
,
Spanswick
C
2000
Blastocyst implantation: the adhesion cascade.
Semin Cell Dev Biol
11
:
77
92

111

Surveyor
GA
,
Gendler
SJ
,
Pemberton
L
,
Das
SK
,
Chakraborty
I
,
Julian
J
,
Pimental
RA
,
Wegner
CC
,
Dey
SK
,
Carson
DD
1995
Expression and steroid hormonal control of Muc-1 in the mouse uterus.
Endocrinology
136
:
3639
3647

112

Hoffman
LH
,
Olson
GE
,
Carson
DD
,
Chilton
BS
1998
Progesterone and implanting blastocysts regulate Muc1 expression in rabbit uterine epithelium.
Endocrinology
139
:
266
271

113

Meseguer
M
,
Aplin
JD
,
Caballero
C
,
Connor
JE
,
Martin
JC
,
Remohi
J
,
Pellicer
A
,
Simon
C
2001
Human endometrial mucin MUC1 is up-regulated by progesterone and down-regulated in vitro by the human blastocyst.
Biol Reprod
64
:
590
601

114

Giancotti
FG
,
Ruoslahti
E
1999
Integrin signaling.
Science
285
:
1028
1032

115

Tabibzadeh
S
1992
Patterns of expression of integrin molecules in human endometrium throughout the menstrual cycle.
Hum Reprod
7
:
876
882

116

Klentzeris
LD
,
Bulmer
JN
,
Trejdosiewicz
LK
,
Morrison
L
,
Cooke
ID
1993
β-1 Integrin cell adhesion molecules in the endometrium of fertile and infertile women.
Hum Reprod
8
:
1223
1230

117

Lessey
BA
,
Damjanovich
L
,
Coutifaris
C
,
Castelbaum
A
,
Albelda
SM
,
Buck
CA
1992
Integrin adhesion molecules in the human endometrium. Correlation with the normal and abnormal menstrual cycle.
J Clin Invest
90
:
188
195

118

Lessey
BA
1994
The use of integrins for the assessment of uterine receptivity.
Fertil Steril
61
:
812
814

119

Lessey
BA
,
Ilesanmi
AO
,
Lessey
MA
,
Riben
M
,
Harris
JE
,
Chwalisz
K
1996
Luminal and glandular endometrial epithelium express integrins differentially throughout the menstrual cycle: implications for implantation, contraception, and infertility.
Am J Reprod Immunol
35
:
195
204

120

Lessey
BA
,
Castelbaum
AJ
,
Sawin
SW
,
Sun
J
1995
Integrins as markers of uterine receptivity in women with primary unexplained infertility.
Fertil Steril
63
:
535
542

121

Illera
MJ
,
Cullinan
E
,
Gui
Y
,
Yuan
L
,
Beyler
SA
,
Lessey
BA
2000
Blockade of the α(v)β(3) integrin adversely affects implantation in the mouse.
Biol Reprod
62
:
1285
1290

122

Sutherland
AE
,
Calarco
PG
,
Damsky
CH
1993
Developmental regulation of integrin expression at the time of implantation in the mouse embryo.
Development
119
:
1175
1186

123

Schultz
JF
,
Armant
DR
1995
β1- And β3-class integrins mediate fibronectin binding activity at the surface of developing mouse peri-implantation blastocysts. Regulation by ligand-induced mobilization of stored receptor.
J Biol Chem
270
:
11522
11531

124

Stephens
LE
,
Sutherland
AE
,
Klimanskaya
IV
,
Andrieux
A
,
Meneses
J
,
Pedersen
RA
,
Damsky
CH
1995
Deletion of β 1 integrins in mice results in inner cell mass failure and peri-implantation lethality.
Genes Dev
9
:
1883
1895

125

Wang
J
,
Mayernik
L
,
Armant
DR
2002
Integrin signaling regulates blastocyst adhesion to fibronectin at implantation: intracellular calcium transients and vesicle trafficking in primary trophoblast cells.
Dev Biol
245
:
270
279

126

Armant
DR
,
Wang
J
,
Liu
Z
2000
Intracellular signaling in the developing blastocyst as a consequence of the maternal-embryonic dialogue.
Semin Reprod Med
18
:
273
287

127

Wang
J
,
Armant
DR
2002
Integrin-mediated adhesion and signaling during blastocyst implantation.
Cells Tissues Organs
172
:
190
201

128

Fassler
R
,
Meyer
M
1995
Consequences of lack of β 1 integrin gene expression in mice.
Genes Dev
9
:
1896
1908

129

Hynes
RO
1996
Targeted mutations in cell adhesion genes: what have we learned from them?
Dev Biol
180
:
402
412

130

Kreidberg
JA
,
Donovan
MJ
,
Goldstein
SL
,
Rennke
H
,
Shepherd
K
,
Jones
RC
,
Jaenisch
R
1996
α3 β1 Integrin has a crucial role in kidney and lung organogenesis.
Development
122
:
3537
3547

131

Gardner
H
,
Kreidberg
J
,
Koteliansky
V
,
Jaenisch
R
1996
Deletion of integrin α1 by homologous recombination permits normal murine development but gives rise to a specific deficit in cell adhesion.
Dev Biol
175
:
301
313

132

Armant
DR
,
Kaplan
HA
,
Lennarz
WJ
1986
Fibronectin and laminin promote in vitro attachment and outgrowth of mouse blastocysts.
Dev Biol
116
:
519
523

133

Carson
DD
,
Tang
JP
,
Gay
S
1988
Collagens support embryo attachment and outgrowth in vitro: effects of the Arg-Gly-Asp sequence.
Dev Biol
127
:
368
375

134

Sutherland
AE
,
Calarco
PG
,
Damsky
CH
1988
Expression and function of cell surface extracellular matrix receptors in mouse blastocyst attachment and outgrowth.
J Cell Biol
106
:
1331
1348

135

Yelian
FD
,
Edgeworth
NA
,
Dong
LJ
,
Chung
AE
,
Armant
DR
1993
Recombinant entactin promotes mouse primary trophoblast cell adhesion and migration through the Arg-Gly-Asp (RGD) recognition sequence.
J Cell Biol
121
:
923
929

136

Wordinger
RJ
,
Jackson
FL
,
Morrill
A
1986
Implantation, deciduoma formation and live births in mast cell-deficient mice (W/Wv).
J Reprod Fertil
77
:
471
476

137

Armant
DR
,
Kameda
S
1994
Mouse trophoblast cell invasion of extracellular matrix purified from endometrial tissue: a model for peri-implantation development.
J Exp Zool
269
:
146
156

138

Wartiovaara
J
,
Leivo
I
,
Vaheri
A
1979
Expression of the cell surface-associated glycoprotein, fibronectin, in the early mouse embryo.
Dev Biol
69
:
247
257

139

Wewer
UM
,
Liotta
LA
,
Jaye
M
,
Ricca
GA
,
Drohan
WN
,
Claysmith
AP
,
Rao
CN
,
Wirth
P
,
Coligan
JE
,
Albrechtsen
R
1986
Altered levels of laminin receptor mRNA in various human carcinoma cells that have different abilities to bind laminin.
Proc Natl Acad Sci USA
83
:
7137
7141

140

Blankenship
TN
,
Enders
AC
,
King
BF
1992
Distribution of laminin, type IV collagen, and fibronectin in the cell columns and trophoblastic shell of early macaque placentas.
Cell Tissue Res
270
:
241
248

141

Armant
DR
,
Kaplan
HA
,
Mover
H
,
Lennarz
WJ
1986
The effect of hexapeptides on attachment and outgrowth of mouse blastocysts cultured in vitro: evidence for the involvement of the cell recognition tripeptide Arg-Gly-Asp.
Proc Natl Acad Sci USA
83
:
6751
6755

142

Hynes
RO
1992
Specificity of cell adhesion in development: the cadherin superfamily.
Curr Opin Genet Dev
2
:
621
624

143

Armant
DR
1991
Cell interactions with laminin and its proteolytic fragments during outgrowth of mouse primary trophoblast cells.
Biol Reprod
45
:
664
672

144

Klaffky
E
,
Williams
R
,
Yao
CC
,
Ziober
B
,
Kramer
R
,
Sutherland
A
2001
Trophoblast-specific expression and function of the integrin α7 subunit in the peri-implantation mouse embryo.
Dev Biol
239
:
161
175

145

Fukuda
MN
,
Sato
T
,
Nakayama
J
,
Klier
G
,
Mikami
M
,
Aoki
D
,
Nozawa
S
1995
Trophinin and tastin, a novel cell adhesion molecule complex with potential involvement in embryo implantation.
Genes Dev
9
:
1199
1210

146

Suzuki
N
,
Nakayama
J
,
Shih
IM
,
Aoki
D
,
Nozawa
S
,
Fukuda
MN
1999
Expression of trophinin, tastin, and bystin by trophoblast and endometrial cells in human placenta.
Biol Reprod
60
:
621
627

147

Suzuki
N
,
Nadano
D
,
Paria
BC
,
Kupriyanov
S
,
Sugihara
K
,
Fukuda
MN
2000
Trophinin expression in the mouse uterus coincides with implantation and is hormonally regulated but not induced by implanting blastocysts.
Endocrinology
141
:
4247
4254

148

Ozawa
M
,
Engel
J
,
Kemler
R
1990
Single amino acid substitutions in one Ca2+ binding site of uvomorulin abolish the adhesive function.
Cell
63
:
1033
1038

149

Ozawa
M
,
Kemler
R
1992
Molecular organization of the uvomorulin-catenin complex.
J Cell Biol
116
:
989
996

150

Larue
L
,
Ohsugi
M
,
Hirchenhain
J
,
Kemler
R
1994
E-cadherin null mutant embryos fail to form a trophectoderm epithelium.
Proc Natl Acad Sci USA
91
:
8263
8267

151

Riethmacher
D
,
Brinkmann
V
,
Birchmeier
C
1995
A targeted mutation in the mouse E-cadherin gene results in defective preimplantation development.
Proc Natl Acad Sci USA
92
:
855
859

152

Paria
BC
,
Zhao
X
,
Das
SK
,
Dey
SK
,
Yoshinaga
K
1999
Zonula occludens-1 and E-cadherin are coordinately expressed in the mouse uterus with the initiation of implantation and decidualization.
Dev Biol
208
:
488
501

153

Wang
X
,
Matsumoto
H
,
Zhao
X
,
Das
SK
,
Paria
BC
2004
Embryonic signals direct formation of tight junctional permeability barrier in the decidualizing stroma during embryo implantation.
J Cell Sci
117
:
53
62

154

Genbacev
OD
,
Prakobphol
A
,
Foulk
RA
,
Krtolica
AR
,
Ilic
D
,
Singer
MS
,
Yang
ZQ
,
Kiessling
LL
,
Rosen
SD
,
Fisher
SJ
2003
Trophoblast L-selectin-mediated adhesion at the maternal-fetal interface.
Science
299
:
405
408

155

Chobotova
K
,
Spyropoulou
I
,
Carver
J
,
Manek
S
,
Heath
JK
,
Gullick
WJ
,
Barlow
DH
,
Sargent
IL
,
Mardon
HJ
2002
Heparin-binding epidermal growth factor and its receptor ErbB4 mediate implantation of the human blastocyst.
Mech Dev
119
:
137
144

156

Hakanson
R
1967
Kinetic properties of mammalian histidine decarboxylase.
Eur J Pharmacol
1
:
42
46

157

Schwartz
JC
1977
Histaminergic mechanisms in brain.
Annu Rev Pharmacol Toxicol
17
:
325
339

158

Marcus
GJ
,
Shelesnyak
MC
1968
Studies on the mechanism of nidation. 33. Coital elevation of uterine histamine content.
Acta Endocrinol (Copenh)
57
:
136
141

159

Shelesnyak
MC
1957
Some experimental studies on the mechanism of ova-implantation in the rat.
Recent Prog Horm Res
13
:
269
–317; discussion 317–322

160

Padilla
L
,
Reinicke
K
,
Montesino
H
,
Villena
F
,
Asencio
H
,
Cruz
M
,
Rudolph
MI
1990
Histamine content and mast cells distribution in mouse uterus: the effect of sexual hormones, gestation and labor.
Cell Mol Biol
36
:
93
100

161

Spaziani
E
1963
Relationship between early vascular responses and growth in the rat uterus: stimulation of cell division by estradiol and vasodilating amines.
Endocrinology
72
:
180
188

162

Marcus
GJ
,
Kracier
PF
,
Shelesnyak
MC
1963
Studies on the mechanism of decidualization. II. The histamine releasing of pyrathiazine.
J Reprod Fertil
5
:
409
415

163

Dey
SK
,
Villanueva
C
,
Chien
SM
,
Crist
RD
1978
The role of histamine in implantation in the rabbit.
J Reprod Fertil
53
:
23
26

164

Inoue
I
,
Yanai
K
,
Kitamura
D
,
Taniuchi
I
,
Kobayashi
T
,
Niimura
K
,
Watanabe
T
,
Watanabe
T
1996
Impaired locomotor activity and exploratory behavior in mice lacking histamine H1 receptors.
Proc Natl Acad Sci USA
93
:
13316
13320

165

Kobayashi
T
,
Tonai
S
,
Ishihara
Y
,
Koga
R
,
Okabe
S
,
Watanabe
T
2000
Abnormal functional and morphological regulation of the gastric mucosa in histamine H2 receptor-deficient mice.
J Clin Invest
105
:
1741
1749

166

Toyota
H
,
Dugovic
C
,
Koehl
M
,
Laposky
AD
,
Weber
C
,
Ngo
K
,
Wu
Y
,
Lee
DH
,
Yanai
K
,
Sakurai
E
,
Watanabe
T
,
Liu
C
,
Chen
J
,
Barbier
AJ
,
Turek
FW
,
Fung
L
,
Lovenberg
TW
2002
Behavioral characterization of mice lacking histamine H(3) receptors.
Mol Pharmacol
62
:
389
397

167

Brandon
JM
,
Wallis
RM
1977
Effect of mepyramine, a histamine H1-, and burimamide, a histamine H2-receptor antagonist, on ovum implantation in the rat.
J Reprod Fertil
50
:
251
254

168

Johnson
DC
,
Dey
SK
1980
Role of histamine in implantation: dexamethasone inhibits estradiol-induced implantation in the rat.
Biol Reprod
22
:
1136
1141

169

Brandon
JM
,
Bibby
MC
1979
A study of changes in uterine mast cells during early pregnancy in the rat.
Biol Reprod
20
:
977
980

170

Salamonsen
LA
,
Jeziorska
M
,
Newlands
GF
,
Dey
SK
,
Woolley
DE
1996
Evidence against a significant role for mast cells in blastocyst implantation in the rat and mouse.
Reprod Fertil Dev
8
:
1157
1164

171

Paria
BC
,
Das
N
,
Das
SK
,
Zhao
X
,
Dileepan
KN
,
Dey
SK
1998
Histidine decarboxylase gene in the mouse uterus is regulated by progesterone and correlates with uterine differentiation for blastocyst implantation.
Endocrinology
139
:
3958
3966

172

Ohtsu
H
,
Tanaka
S
,
Terui
T
,
Hori
Y
,
Makabe
K
,
Pejler
G
,
Tchougounova
E
,
Hellman
L
,
Gertsenstein
M
,
Hirasawa
N
,
Sakurai
E
,
Buzas
E
,
Kovacs
P
,
Csaba
G
,
Kittel
A
,
Okada
M
,
Hara
M
,
Mar
L
,
Numayama
T
,
Ishigaki
S
,
Ohuchi
K
,
Ichikawa
A
,
Falus
A
,
Watanabe
T
,
Nagy
A
2001
Mice lacking histidine decarboxylase exhibit abnormal mast cells.
FEBS Lett
502
:
53
56

173

Smith
WL
,
Dewitt
DL
1996
Prostaglandin endoperoxide H synthases-1 and -2.
Adv Immunol
62
:
167
215

174

Forman
BM
,
Chen
J
,
Evans
RM
1997
Hypolipidemic drugs, polyunsaturated fatty acids, and eicosanoids are ligands for peroxisome proliferator-activated receptors α and δ.
Proc Natl Acad Sci USA
94
:
4312
4317

175

Forman
BM
,
Tontonoz
P
,
Chen
J
,
Brun
RP
,
Spiegelman
BM
,
Evans
RM
1995
15-Deoxy-δ 12,14-prostaglandin J2 is a ligand for the adipocyte determination factor PPAR γ.
Cell
83
:
803
812

176

Kliewer
SA
,
Umesono
K
,
Noonan
DJ
,
Heyman
RA
,
Evans
RM
1992
Convergence of 9-cis retinoic acid and peroxisome proliferator signalling pathways through heterodimer formation of their receptors.
Nature
358
:
771
774

177

Kliewer
SA
,
Forman
BM
,
Blumberg
B
,
Ong
ES
,
Borgmeyer
U
,
Mangelsdorf
DJ
,
Umesono
K
,
Evans
RM
1994
Differential expression and activation of a family of murine peroxisome proliferator-activated receptors.
Proc Natl Acad Sci USA
91
:
7355
7359

178

Copeland
RA
,
Williams
JM
,
Giannaras
J
,
Nurnberg
S
,
Covington
M
,
Pinto
D
,
Pick
S
,
Trzaskos
JM
1994
Mechanism of selective inhibition of the inducible isoform of prostaglandin G/H synthase.
Proc Natl Acad Sci USA
91
:
11202
11206

179

Kurumbail
RG
,
Stevens
AM
,
Gierse
JK
,
McDonald
JJ
,
Stegeman
RA
,
Pak
JY
,
Gildehaus
D
,
Miyashiro
JM
,
Penning
TD
,
Seibert
K
,
Isakson
PC
,
Stallings
WC
1996
Structural basis for selective inhibition of cyclooxygenase-2 by anti-inflammatory agents.
Nature
384
:
644
648

180

DuBois
RN
,
Radhika
A
,
Reddy
BS
,
Entingh
AJ
1996
Increased cyclooxygenase-2 levels in carcinogen-induced rat colonic tumors.
Gastroenterology
110
:
1259
1262

181

Oshima
M
,
Dinchuk
JE
,
Kargman
SL
,
Oshima
H
,
Hancock
B
,
Kwong
E
,
Trzaskos
JM
,
Evans
JF
,
Taketo
MM
1996
Suppression of intestinal polyposis in Apc δ716 knockout mice by inhibition of cyclooxygenase 2 (COX-2).
Cell
87
:
803
809

182

Espey
LL
1994
Current status of the hypothesis that mammalian ovulation is comparable to an inflammatory reaction.
Biol Reprod
50
:
233
238

183

McMaster
MT
,
Dey
SK
,
Andrews
GK
1993
Association of monocytes and neutrophils with early events of blastocyst implantation in mice.
J Reprod Fertil
99
:
561
569

184

Sirois
J
1994
Induction of prostaglandin endoperoxide synthase-2 by human chorionic gonadotropin in bovine preovulatory follicles in vivo.
Endocrinology
135
:
841
848

185

Chakraborty
I
,
Das
SK
,
Wang
J
,
Dey
SK
1996
Developmental expression of the cyclo-oxygenase-1 and cyclo-oxygenase-2 genes in the peri-implantation mouse uterus and their differential regulation by the blastocyst and ovarian steroids.
J Mol Endocrinol
16
:
107
122

186

Dinchuk
JE
,
Car
BD
,
Focht
RJ
,
Johnston
JJ
,
Jaffee
BD
,
Covington
MB
,
Contel
NR
,
Eng
VM
,
Collins
RJ
,
Czerniak
PM
1995
Renal abnormalities and an altered inflammatory response in mice lacking cyclooxygenase II.
Nature
378
:
406
409

187

Langenbach
R
,
Morham
SG
,
Tiano
HF
,
Loftin
CD
,
Ghanayem
BI
,
Chulada
PC
,
Mahler
JF
,
Lee
CA
,
Goulding
EH
,
Kluckman
KD
1995
Prostaglandin synthase 1 gene disruption in mice reduces arachidonic acid-induced inflammation and indomethacininduced gastric ulceration.
Cell
83
:
483
492

188

Lim
H
,
Paria
BC
,
Das
SK
,
Dinchuk
JE
,
Langenbach
R
,
Trzaskos
JM
,
Dey
SK
1997
Multiple female reproductive failures in cyclooxygenase 2-deficient mice.
Cell
91
:
197
208

189

Reese
J
,
Brown
N
,
Paria
BC
,
Morrow
J
,
Dey
SK
1999
COX-2 compensation in the uterus of COX-1 deficient mice during the pre-implantation period.
Mol Cell Endocrinol
150
:
23
31

190

Lim
H
,
Gupta
RA
,
Ma
WG
,
Paria
BC
,
Moller
DE
,
Morrow
JD
,
DuBois
RN
,
Trzaskos
JM
,
Dey
SK
1999
Cyclo-oxygenase-2derived prostacyclin mediates embryo implantation in the mouse via PPARδ.
Genes Dev
13
:
1561
1574

191

Clark
JD
,
Schievella
AR
,
Nalefski
EA
,
Lin
LL
1995
Cytosolic phospholipase A2.
J Lipid Mediat Cell Signal
12
:
83
117

192

Bonventre
JV
,
Huang
Z
,
Taheri
MR
,
Leary
E
,
Li
E
,
Moskowitz
MA
,
Sapirstein
A
1997
Reduced fertility and postischaemic brain injury in mice deficient in cytosolic phospholipase A2.
Nature
390
:
622
625

193

Uozumi
N
,
Kume
K
,
Nagase
T
,
Nakatani
N
,
Ishii
S
,
Tashiro
F
,
Komagata
Y
,
Maki
K
,
Ikuta
K
,
Ouchi
Y
,
Miyazaki
J
,
Shimizu
T
1997
Role of cytosolic phospholipase A2 in allergic response and parturition.
Nature
390
:
618
622

194

Cheng
JG
,
Stewart
CL
2003
Loss of cyclooxygenase-2 retards decidual growth but does not inhibit embryo implantation or development to term.
Biol Reprod
68
:
401
404

195

Charpigny
G
,
Reinaud
P
,
Tamby
JP
,
Creminon
C
,
Guillomot
M
1997
Cyclooxygenase-2 unlike cyclooxygenase-1 is highly expressed in ovine embryos during the implantation period.
Biol Reprod
57
:
1032
1040

196

Das
SK
,
Wang
J
,
Dey
SK
,
Mead
RA
1999
Spatiotemporal expression of cyclooxygenase 1 and cyclooxygenase 2 during delayed implantation and the periimplantation period in the Western spotted skunk.
Biol Reprod
60
:
893
899

197

Song
JH
,
Sirois
J
,
Houde
A
,
Murphy
BD
1998
Cloning, developmental expression, and immunohistochemistry of cyclooxygenase 2 in the endometrium during embryo implantation and gestation in the mink (Mustela vison).
Endocrinology
139
:
3629
3636

198

Kim
JJ
,
Wang
J
,
Bambra
C
,
Das
SK
,
Dey
SK
,
Fazleabas
AT
1999
Expression of cyclooxygenase-1 and -2 in the baboon endometrium during the menstrual cycle and pregnancy.
Endocrinology
140
:
2672
2678

199

Critchley
HO
,
Jones
RL
,
Lea
RG
,
Drudy
TA
,
Kelly
RW
,
Williams
AR
,
Baird
DT
1999
Role of inflammatory mediators in human endometrium during progesterone withdrawal and early pregnancy.
J Clin Endocrinol Metab
84
:
240
248

200

Marions
L
,
Danielsson
KG
1999
Expression of cyclo-oxygenase in human endometrium during the implantation period.
Mol Hum Reprod
5
:
961
965

201

Wang
H
,
Ma
WG
,
Tejada
L
,
Zhang
H
,
Morrow
J
,
Das
SK
,
Dey
SK
2004
Rescue of female infertility from the loss of cyclooxygenase-2 by compensatory upregulation of cyclooxygenase-1 is a function of genetic background.
J Biol Chem
279
:
10649
10658

202

Lim
H
,
Dey
SK
1999
PPAR δ functions as a prostacyclin receptor in blastocyst implantation.
Trends Endocrinol Metab
11
:
137
142

203

Yang
ZM
,
Das
SK
,
Wang
J
,
Sugimoto
Y
,
Ichikawa
A
,
Dey
SK
1997
Potential sites of prostaglandin actions in the periimplantation mouse uterus: differential expression and regulation of prostaglandin receptor genes.
Biol Reprod
56
:
368
379

204

Ni
H
,
Sun
T
,
Ma
XH
,
Yang
ZM
2003
Expression and regulation of cytosolic prostaglandin E synthase in mouse uterus during the peri-implantation period.
Biol Reprod
68
:
744
750

205

Ni
H
,
Sun
T
,
Ding
NZ
,
Ma
XH
,
Yang
ZM
2002
Differential expression of microsomal prostaglandin E synthase at implantation sites and in decidual cells of mouse uterus.
Biol Reprod
67
:
351
358

206

Kennedy
TG
1985
Evidence for the involvement of prostaglandins throughout the decidual cell reaction in the rat.
Biol Reprod
33
:
140
146

207

Racowsky
C
,
Biggers
JD
1983
Are blastocyst prostaglandins produced endogenously?
Biol Reprod
29
:
379
388

208

Reese
J
,
Paria
BC
,
Brown
N
,
Zhao
X
,
Morrow
JD
,
Dey
SK
2000
Coordinated regulation of fetal and maternal prostaglandins directs successful birth and postnatal adaptation in the mouse.
Proc Natl Acad Sci USA
97
:
9759
9764

209

Cross
JC
,
Werb
Z
,
Fisher
SJ
1994
Implantation and the placenta: key pieces of the development puzzle.
Science
266
:
1508
1518

210

Tazuke
SI
,
Giudice
LC
1996
Growth factors and cytokines in endometrium, embryonic development, and maternal: embryonic interactions.
Semin Reprod Endocrinol
14
:
231
245

211

Das
SK
,
Das
N
,
Wang
J
,
Lim
H
,
Schryver
B
,
Plowman
GD
,
Dey
SK
1997
Expression of betacellulin and epiregulin genes in the mouse uterus temporally by the blastocyst solely at the site of its apposition is coincident with the “window” of implantation.
Dev Biol
190
:
178
190

212

Das
SK
,
Chakraborty
I
,
Paria
BC
,
Wang
XN
,
Plowman
G
,
Dey
SK
1995
Amphiregulin is an implantation-specific and progesterone-regulated gene in the mouse uterus.
Mol Endocrinol
9
:
691
705

213

Luetteke
NC
,
Qiu
TH
,
Peiffer
RL
,
Oliver
P
,
Smithies
O
,
Lee
DC
1993
TGFα deficiency results in hair follicle and eye abnormalities in targeted and waved-1 mice.
Cell
73
:
263
278

214

Luetteke
NC
,
Qiu
TH
,
Fenton
SE
,
Troyer
KL
,
Riedel
RF
,
Chang
A
,
Lee
DC
1999
Targeted inactivation of the EGF and amphiregulin genes reveals distinct roles for EGF receptor ligands in mouse mammary gland development.
Development
126
:
2739
2750

215

Olayioye
MA
,
Neve
RM
,
Lane
HA
,
Hynes
NE
2000
The ErbB signaling network: receptor heterodimerization in development and cancer.
EMBO J
19
:
3159
3167

216

Harvey
MB
,
Leco
KJ
,
Arcellana
P
,
Zhang
X
,
Edwards
DR
,
Schultz
GA
1995
Roles of growth factors during peri-implantation development.
Hum Reprod
10
:
712
718

217

Rappolee
DA
,
Brenner
CA
,
Schultz
R
,
Mark
D
,
Werb
Z
1988
Developmental expression of PDGF, TGF-α, and TGF-β genes in preimplantation mouse embryos.
Science
241
:
1823
1825

218

Paria
BC
,
Elenius
K
,
Klagsbrun
M
,
Dey
SK
1999
Heparin-binding EGF-like growth factor interacts with mouse blastocysts independently of ErbB1: a possible role for heparan sulfate proteoglycans and ErbB4 in blastocyst implantation.
Development
126
:
1997
2005

219

Das
SK
,
Tsukamura
H
,
Paria
BC
,
Andrews
GK
,
Dey
SK
1994
Differential expression of epidermal growth factor receptor (EGF-R) gene and regulation of EGF-R bioactivity by progesterone and estrogen in the adult mouse uterus.
Endocrinology
134
:
971
981

220

Iwamoto
R
,
Yamazaki
S
,
Asakura
M
,
Takashima
S
,
Hasuwa
H
,
Miyado
K
,
Adachi
S
,
Kitakaze
M
,
Hashimoto
K
,
Raab
G
,
Nanba
D
,
Higashiyama
S
,
Hori
M
,
Klagsbrun
M
,
Mekada
E
2003
Heparin-binding EGF-like growth factor and ErbB signaling is essential for heart function.
Proc Natl Acad Sci USA
100
:
3221
3226

221

Gassmann
M
,
Casagranda
F
,
Orioli
D
,
Simon
H
,
Lai
C
,
Klein
R
,
Lemke
G
1995
Aberrant neural and cardiac development in mice lacking the ErbB4 neuregulin receptor.
Nature
378
:
390
394

222

Threadgill
DW
,
Dlugosz
AA
,
Hansen
LA
,
Tennenbaum
T
,
Lichti
U
,
Yee
D
,
LaMantia
C
,
Mourton
T
,
Herrup
K
,
Harris
RC
1995
Targeted disruption of mouse EGF receptor: effect of genetic background on mutant phenotype.
Science
269
:
230
234

223

Yoo
HJ
,
Barlow
DH
,
Mardon
HJ
1997
Temporal and spatial regulation of expression of heparin-binding epidermal growth factor-like growth factor in the human endometrium: a possible role in blastocyst implantation.
Dev Genet
21
:
102
108

224

Leach
RE
,
Khalifa
R
,
Ramirez
ND
,
Das
SK
,
Wang
J
,
Dey
SK
,
Romero
R
,
Armant
DR
1999
Multiple roles for heparin-binding epidermal growth factor-like growth factor are suggested by its cell-specific expression during the human endometrial cycle and early placentation.
J Clin Endocrinol Metab
84
:
3355
3363

225

Martin
KL
,
Barlow
DH
,
Sargent
IL
1998
Heparin-binding epidermal growth factor significantly improves human blastocyst development and hatching in serum-free medium.
Hum Reprod
13
:
1645
1652

226

Chard
T
1995
Cytokines in implantation.
Hum Reprod Update
1
:
385
396

227

Pollard
JW
,
Hunt
JS
,
Wiktor
J
,
Stanley
ER
1991
A pregnancy defect in the osteopetrotic (op/op) mouse demonstrates the requirement for CSF-1 in female fertility.
Dev Biol
148
:
273
283

228

Stewart
CL
,
Kaspar
P
,
Brunet
LJ
,
Bhatt
H
,
Gadi
I
,
Kontgen
F
,
Abbondanzo
SJ
1992
Blastocyst implantation depends on maternal expression of leukaemia inhibitory factor.
Nature
359
:
76
79

229

Robb
L
,
Li
R
,
Hartley
L
,
Nandurkar
HH
,
Koentgen
F
,
Begley
CG
1998
Infertility in female mice lacking the receptor for interleukin 11 is due to a defective uterine response to implantation.
Nat Med
4
:
303
308

230

Escary
JL
,
Perreau
J
,
Dumenil
D
,
Ezine
S
,
Brulet
P
1993
Leukaemia inhibitory factor is necessary for maintenance of haematopoietic stem cells and thymocyte stimulation.
Nature
363
:
361
364

231

Kishimoto
T
,
Taga
T
,
Akira
S
1994
Cytokine signal transduction.
Cell
76
:
253
262

232

Song
H
,
Lim
H
,
Das
SK
,
Paria
BC
,
Dey
SK
2000
Dysregulation of EGF family of growth factors and COX-2 in the uterus during the preattachment and attachment reactions of the blastocyst with the luminal epithelium correlates with implantation failure in LIF-deficient mice.
Mol Endocrinol
14
:
1147
1161

233

Rathjen
PD
,
Toth
S
,
Willis
A
,
Heath
JK
,
Smith
AG
1990
Differentiation inhibiting activity is produced in matrix-associated and diffusible forms that are generated by alternate promoter usage.
Cell
62
:
1105
1114

234

Narazaki
M
,
Witthuhn
BA
,
Yoshida
K
,
Silvennoinen
O
,
Yasukawa
K
,
Ihle
JN
,
Kishimoto
T
,
Taga
T
1994
Activation of JAK2 kinase mediated by the interleukin 6 signal transducer gp130.
Proc Natl Acad Sci USA
91
:
2285
2289

235

Layton
MJ
,
Lock
P
,
Metcalf
D
,
Nicola
NA
1994
Cross-species receptor binding characteristics of human and mouse leukemia inhibitory factor suggest a complex binding interaction.
J Biol Chem
269
:
17048
17055

236

Bhatt
H
,
Brunet
LJ
,
Stewart
CL
1991
Uterine expression of leukemia inhibitory factor coincides with the onset of blastocyst implantation.
Proc Natl Acad Sci USA
88
:
11408
11412

237

Ernst
M
,
Inglese
M
,
Waring
P
,
Campbell
IK
,
Bao
S
,
Clay
FJ
,
Alexander
WS
,
Wicks
IP
,
Tarlinton
DM
,
Novak
U
,
Heath
JK
,
Dunn
AR
2001
Defective gp130-mediated signal transducer and activator of transcription (STAT) signaling results in degenerative joint disease, gastrointestinal ulceration, and failure of uterine implantation.
J Exp Med
194
:
189
203

238

Fry
RC
1992
The effect of leukaemia inhibitory factor (LIF) on embryogenesis.
Reprod Fertil Dev
4
:
449
458

239

Dunglison
GF
,
Barlow
DH
,
Sargent
IL
1996
Leukaemia inhibitory factor significantly enhances the blastocyst formation rates of human embryos cultured in serum-free medium.
Hum Reprod
11
:
191
196

240

Nichols
J
,
Davidson
D
,
Taga
T
,
Yoshida
K
,
Chambers
I
,
Smith
A
1996
Complementary tissue-specific expression of LIF and LIF-receptor mRNAs in early mouse embryogenesis.
Mech Dev
57
:
123
131

241

Ware
CB
,
Horowitz
MC
,
Renshaw
BR
,
Hunt
JS
,
Liggitt
D
,
Koblar
SA
,
Gliniak
BC
,
McKenna
HJ
,
Papayannopoulou
T
,
Thoma
B
1995
Targeted disruption of the low-affinity leukemia inhibitory factor receptor gene causes placental, skeletal, neural and metabolic defects and results in perinatal death.
Development
121
:
1283
1299

242

Yoshida
K
,
Taga
T
,
Saito
M
,
Suematsu
S
,
Kumanogoh
A
,
Tanaka
T
,
Fujiwara
H
,
Hirata
M
,
Yamagami
T
,
Nakahata
T
,
Hirabayashi
T
,
Yoneda
Y
,
Tanaka
K
,
Wang
WZ
,
Mori
C
,
Shiota
K
,
Yoshida
N
,
Kishimoto
T
1996
Targeted disruption of gp130, a common signal transducer for the interleukin 6 family of cytokines, leads to myocardial and hematological disorders.
Proc Natl Acad Sci USA
93
:
407
411

243

Shen
MM
,
Leder
P
1992
Leukemia inhibitory factor is expressed by the preimplantation uterus and selectively blocks primitive ectoderm formation in vitro.
Proc Natl Acad Sci USA
89
:
8240
8244

244

Yang
ZM
,
Chen
DB
,
Le
SP
,
Harper
MJ
1996
Differential hormonal regulation of leukemia inhibitory factor (LIF) in rabbit and mouse uterus.
Mol Reprod Dev
43
:
470
476

245

Arici
A
,
Engin
O
,
Attar
E
,
Olive
DL
1995
Modulation of leukemia inhibitory factor gene expression and protein biosynthesis in human endometrium.
J Clin Endocrinol Metab
80
:
1908
1915

246

Hambartsoumian
E
1998
Endometrial leukemia inhibitory factor (LIF) as a possible cause of unexplained infertility and multiple failures of implantation.
Am J Reprod Immunol
39
:
137
143

247

Krumlauf
R
1994
Hox genes in vertebrate development.
Cell
78
:
191
201

248

Benson
GV
,
Lim
H
,
Paria
BC
,
Satokata
I
,
Dey
SK
,
Maas
RL
1996
Mechanisms of reduced fertility in Hoxa-10 mutant mice: uterine homeosis and loss of maternal Hoxa-10 expression.
Development
122
:
2687
2696

249

Hsieh
L
,
Witte
DP
,
Weinstein
M
,
Branford
W
,
Li
H
,
Small
K
,
Potter
SS
1995
Hoxa 11 structure, extensive antisense transcription, and function in male and female fertility.
Development
121
:
1373
1385

250

Gendron
RL
,
Paradis
H
,
Hsieh
L
,
Lee
DW
,
Potter
SS
,
Markoff
E
1997
Abnormal uterine stromal and glandular function associated with maternal reproductive defects in Hoxa-11 null mice.
Biol Reprod
56
:
1097
1105

251

Lim
H
,
Ma
L
,
Ma
WG
,
Maas
RL
,
Dey
SK
1999
Hoxa-10 regulates uterine stromal cell responsiveness to progesterone during implantation and decidualization in the mouse.
Mol Endocrinol
13
:
1005
1017

252

Das
SK
,
Lim
H
,
Paria
BC
,
Dey
SK
1999
Cyclin D3 in the mouse uterus is associated with the decidualization process during early pregnancy.
J Mol Endocrinol
22
:
91
101

253

Yao
MW
,
Lim
H
,
Schust
DJ
,
Choe
SE
,
Farago
A
,
Ding
Y
,
Michaud
S
,
Church
GM
,
Maas
RL
2003
Gene expression profiling reveals progesterone-mediated cell cycle and immunoregulatory roles of hoxa-10 in the preimplantation uterus.
Mol Endocrinol
17
:
610
627

254

Taylor
HS
,
Arici
A
,
Olive
D
,
Igarashi
P
1998
HOXA10 is expressed in response to sex steroids at the time of implantation in the human endometrium.
J Clin Invest
101
:
1379
1384

255

Wang
W
,
Van
D
,
Lufkin
T
1998
Inner ear and maternal reproductive defects in mice lacking the Hmx3 homeobox gene.
Development
125
:
621
634

256

Desvergne
B
,
Wahli
W
1999
Peroxisome proliferator-activated receptors: nuclear control of metabolism.
Endocr Rev
20
:
649
688

257

Ding
NZ
,
Teng
CB
,
Ma
H
,
Ni
H
,
Ma
XH
,
Xu
LB
,
Yang
ZM
2003
Peroxisome proliferator-activated receptor δ expression and regulation in mouse uterus during embryo implantation and decidualization.
Mol Reprod Dev
66
:
218
224

258

Lim
H
,
Dey
SK
2002
A novel pathway of prostacyclin signaling-hanging out with nuclear receptors.
Endocrinology
143
:
3207
3210

259

Barak
Y
,
Liao
D
,
He
W
,
Ong
ES
,
Nelson
MC
,
Olefsky
JM
,
Boland
R
,
Evans
RM
2002
Effects of peroxisome proliferator-activated receptor δ on placentation, adiposity, and colorectal cancer.
Proc Natl Acad Sci USA
99
:
303
308

260

Peters
JM
,
Lee
SS
,
Li
W
,
Ward
JM
,
Gavrilova
O
,
Everett
C
,
Reitman
ML
,
Hudson
LD
,
Gonzalez
FJ
2000
Growth, adipose, brain, and skin alterations resulting from targeted disruption of the mouse peroxisome proliferator-activated receptor β(δ).
Mol Cell Biol
20
:
5119
5128

261

Michalik
L
,
Desvergne
B
,
Tan
NS
,
Basu-Modak
S
,
Escher
P
,
Rieusset
J
,
Peters
JM
,
Kaya
G
,
Gonzalez
FJ
,
Zakany
J
,
Metzger
D
,
Chambon
P
,
Duboule
D
,
Wahli
W
2001
Impaired skin wound healing in peroxisome proliferator-activated receptor (PPAR)α and PPARβ mutant mice.
J Cell Biol
154
:
799
814

262

Berger
J
,
Moller
DE
2002
The mechanisms of action of PPARs.
Annu Rev Med
53
:
409
435

263

Barak
Y
,
Nelson
MC
,
Ong
ES
,
Jones
YZ
,
Ruiz
L
,
Chien
KR
,
Koder
A
,
Evans
RM
1999
PPARγ is required for placental, cardiac, and adipose tissue development.
Mol Cell
4
:
585
595

264

Cui
Y
,
Miyoshi
K
,
Claudio
E
,
Siebenlist
UK
,
Gonzalez
FJ
,
Flaws
J
,
Wagner
KU
,
Hennighausen
L
2002
Loss of the peroxisome proliferation-activated receptor γ (PPARγ) does not affect mammary development and propensity for tumor formation but leads to reduced fertility.
J Biol Chem
277
:
17830
17835

265

Muerhoff
AS
,
Griffin
KJ
,
Johnson
EF
1992
The peroxisome proliferator-activated receptor mediates the induction of CYP4A6, a cytochrome P450 fatty acid ω-hydroxylase, by clofibric acid.
J Biol Chem
267
:
19051
19053

266

Tugwood
JD
,
Issemann
I
,
Anderson
RG
,
Bundell
KR
,
McPheat
WL
,
Green
S
1992
The mouse peroxisome proliferator activated receptor recognizes a response element in the 5′ flanking sequence of the rat acyl CoA oxidase gene.
EMBO J
11
:
433
439

267

Robyr
D
,
Wolffe
AP
,
Wahli
W
2000
Nuclear hormone receptor coregulators in action: diversity for shared tasks.
Mol Endocrinol
14
:
329
347

268

McKenna
NJ
,
Lanz
RB
,
O’Malley
BW
1999
Nuclear receptor coregulators: cellular and molecular biology.
Endocr Rev
20
:
321
344

269

Yao
TP
,
Oh
SP
,
Fuchs
M
,
Zhou
ND
,
Ch’ng
LE
,
Newsome
D
,
Bronson
RT
,
Li
E
,
Livingston
DM
,
Eckner
R
1998
Gene dosage-dependent embryonic development and proliferation defects in mice lacking the transcriptional integrator p300.
Cell
93
:
361
372

270

Hong
H
,
Kohli
K
,
Trivedi
A
,
Johnson
DL
,
Stallcup
MR
1996
GRIP1, a novel mouse protein that serves as a transcriptional coactivator in yeast for the hormone binding domains of steroid receptors.
Proc Natl Acad Sci USA
93
:
4948
4952

271

Anzick
SL
,
Kononen
J
,
Walker
RL
,
Azorsa
DO
,
Tanner
MM
,
Guan
XY
,
Sauter
G
,
Kallioniemi
OP
,
Trent
JM
,
Meltzer
PS
1997
AIB1, a steroid receptor coactivator amplified in breast and ovarian cancer.
Science
277
:
965
968

272

Xu
J
,
Qiu
Y
,
DeMayo
FJ
,
Tsai
SY
,
Tsai
MJ
,
O’Malley
BW
1998
Partial hormone resistance in mice with disruption of the steroid receptor coactivator-1 (SRC-1) gene.
Science
279
:
1922
1925

273

Gehin
M
,
Mark
M
,
Dennefeld
C
,
Dierich
A
,
Gronemeyer
H
,
Chambon
P
2002
The function of TIF2/GRIP1 in mouse reproduction is distinct from those of SRC-1 and p/CIP.
Mol Cell Biol
22
:
5923
5937

274

Xu
J
,
Liao
L
,
Ning
G
,
Yoshida
K
,
Deng
C
,
O’Malley
BW
2000
The steroid receptor coactivator SRC-3 (p/CIP/RAC3/AIB1/ACTR/TRAM-1) is required for normal growth, puberty, female reproductive function, and mammary gland development.
Proc Natl Acad Sci USA
97
:
6379
6384

275

Yuan
CX
,
Ito
M
,
Fondell
JD
,
Fu
ZY
,
Roeder
RG
1998
The TRAP220 component of a thyroid hormone receptor-associated protein (TRAP) coactivator complex interacts directly with nuclear receptors in a ligand-dependent fashion.
Proc Natl Acad Sci USA
95
:
7939
7944

276

Rachez
C
,
Lemon
BD
,
Suldan
Z
,
Bromleigh
V
,
Gamble
M
,
Naar
AM
,
Erdjument
B
,
Tempst
P
,
Freedman
LP
1999
Ligand-dependent transcription activation by nuclear receptors requires the DRIP complex.
Nature
398
:
824
828

277

Zhu
Y
,
Qi
C
,
Jain
S
,
Rao
MS
,
Reddy
JK
1997
Isolation and characterization of PBP, a protein that interacts with peroxisome proliferator-activated receptor.
J Biol Chem
272
:
25500
25506

278

Zhu
Y
,
Qi
C
,
Jain
S
,
Le
B
,
Espinosa
R
,
Atkins
GB
,
Lazar
MA
,
Yeldandi
AV
,
Rao
MS
,
Reddy
JK
1999
Amplification and overexpression of peroxisome proliferator-activated receptor binding protein (PBP/PPARBP) gene in breast cancer.
Proc Natl Acad Sci USA
96
:
10848
10853

279

Zhu
Y
,
Qi
C
,
Jia
Y
,
Nye
JS
,
Rao
MS
,
Reddy
JK
2000
Deletion of PBP/PPARBP, the gene for nuclear receptor coactivator peroxisome proliferator-activated receptor-binding protein, results in embryonic lethality.
J Biol Chem
275
:
14779
14782

280

Ito
M
,
Yuan
CX
,
Okano
HJ
,
Darnell
RB
,
Roeder
RG
2000
Involvement of the TRAP220 component of the TRAP/SMCC coactivator complex in embryonic development and thyroid hormone action.
Mol Cell
5
:
683
693

281

Zhu
Y
,
Kan
L
,
Qi
C
,
Kanwar
YS
,
Yeldandi
AV
,
Rao
MS
,
Reddy
JK
2000
Isolation and characterization of peroxisome proliferator-activated receptor (PPAR) interacting protein (PRIP) as a coactivator for PPAR.
J Biol Chem
275
:
13510
13516

282

Caira
F
,
Antonson
P
,
Pelto
H
,
Treuter
E
,
Gustafsson
JA
2000
Cloning and characterization of RAP250, a novel nuclear receptor coactivator.
J Biol Chem
275
:
5308
5317

283

Kuang
SQ
,
Liao
L
,
Zhang
H
,
Pereira
FA
,
Yuan
Y
,
DeMayo
FJ
,
Ko
L
,
Xu
J
2002
Deletion of the cancer-amplified coactivator AIB3 results in defective placentation and embryonic lethality.
J Biol Chem
277
:
45356
45360

284

Antonson
P
,
Schuster
GU
,
Wang
L
,
Rozell
B
,
Holter
E
,
Flodby
P
,
Treuter
E
,
Holmgren
L
,
Gustafsson
JA
2003
Inactivation of the nuclear receptor coactivator RAP250 in mice results in placental vascular dysfunction.
Mol Cell Biol
23
:
1260
1268

285

Treuter
E
,
Albrektsen
T
,
Johansson
L
,
Leers
J
,
Gustafsson
JA
1998
A regulatory role for RIP140 in nuclear receptor activation.
Mol Endocrinol
12
:
864
881

286

Lee
CH
,
Chinpaisal
C
,
Wei
LN
1998
Cloning and characterization of mouse RIP140, a corepressor for nuclear orphan receptor TR2.
Mol Cell Biol
18
:
6745
6755

287

White
R
,
Leonardsson
G
,
Rosewell
I
,
Ann
J
,
Milligan
S
,
Parker
M
2000
The nuclear receptor co-repressor nrip1 (RIP140) is essential for female fertility.
Nat Med
6
:
1368
1374

288

Leonardsson
G
,
Jacobs
MA
,
White
R
,
Jeffery
R
,
Poulsom
R
,
Milligan
S
,
Parker
M
2002
Embryo transfer experiments and ovarian transplantation identify the ovary as the only site in which nuclear receptor interacting protein 1/RIP140 action is crucial for female fertility.
Endocrinology
143
:
700
707

289

Sachs
L
,
Shelesnyak
MC
1955
The development and suppression of polyploidy in the developing and suppressed deciduoma in the rat.
J Endocrinol
12
:
146
151

290

Leroy
F
,
Bogaert
C
,
Van
H
,
Delcroix
C
1974
Cytophotometric and autoradiographic evaluation of cell kinetics in decidual growth in rats.
J Reprod Fertil
38
:
441
449

291

Ansell
JD
,
Barlow
PW
,
McLaren
A
1974
Binucleate and polyploid cells in the decidua of the mouse.
J Embryol Exp Morphol
31
:
223
227

292

Sherr
CJ
1993
Mammalian G1 cyclins.
Cell
73
:
1059
1065

293

Riley
DJ
,
Lee
EY
,
Lee
WH
1994
The retinoblastoma protein: more than a tumor suppressor.
Annu Rev Cell Biol
10
:
1
29

294

Resnitzky
D
,
Gossen
M
,
Bujard
H
,
Reed
SI
1994
Acceleration of the G1/S phase transition by expression of cyclins D1 and E with an inducible system.
Mol Cell Biol
14
:
1669
1679

295

Sherr
CJ
,
Roberts
JM
1999
CDK inhibitors: positive and negative regulators of G1-phase progression.
Genes Dev
13
:
1501
1512

296

Shiozawa
T
,
Li
SF
,
Nakayama
K
,
Nikaido
T
,
Fujii
S
1996
Relationship between the expression of cyclins/cyclin-dependent kinases and sex-steroid receptors/Ki67 in normal human endometrial glands and stroma during the menstrual cycle.
Mol Hum Reprod
2
:
745
752

297

Shiozawa
T
,
Nikaido
T
,
Nakayama
K
,
Lu
X
,
Fujii
S
1998
Involvement of cyclin-dependent kinase inhibitor p27Kip1 in growth inhibition of endometrium in the secretory phase and of hyperplastic endometrium treated with progesterone.
Mol Hum Reprod
4
:
899
905

298

Geum
D
,
Sun
W
,
Paik
SK
,
Lee
CC
,
Kim
K
1997
Estrogen-induced cyclin D1 and D3 gene expressions during mouse uterine cell proliferation in vivo: differential induction mechanism of cyclin D1 and D3.
Mol Reprod Dev
46
:
450
458

299

Altucci
L
,
Addeo
R
,
Cicatiello
L
,
Germano
D
,
Pacilio
C
,
Battista
T
,
Cancemi
M
,
Petrizzi
VB
,
Bresciani
F
,
Weisz
A
1997
Estrogen induces early and timed activation of cyclin-dependent kinases 4, 5, and 6 and increases cyclin messenger ribonucleic acid expression in rat uterus.
Endocrinology
138
:
978
984

300

Prall
OW
,
Sarcevic
B
,
Musgrove
EA
,
Watts
CK
,
Sutherland
RL
1997
Estrogen-induced activation of Cdk4 and Cdk2 during G1-S phase progression is accompanied by increased cyclin D1 expression and decreased cyclin-dependent kinase inhibitor association with cyclin E-Cdk2.
J Biol Chem
272
:
10882
10894

301

Musgrove
EA
,
Swarbrick
A
,
Lee
CS
,
Cornish
AL
,
Sutherland
RL
1998
Mechanisms of cyclin-dependent kinase inactivation by progestins.
Mol Cell Biol
18
:
1812
1825

302

Jones
SR
,
Kimler
BF
,
Justice
WM
,
Rider
V
2000
Transit of normal rat uterine stromal cells through G1 phase of the cell cycle requires progesterone-growth factor interactions.
Endocrinology
141
:
637
648

303

Bamberger
A
,
Sudahl
S
,
Bamberger
CM
,
Schulte
HM
,
Loning
T
1999
Expression patterns of the cell-cycle inhibitor p27 and the cell-cycle promoter cyclin E in the human placenta throughout gestation: implications for the control of proliferation.
Placenta
20
:
401
406
.

304

MacAuley
A
,
Cross
JC
,
Werb
Z
1998
Reprogramming the cell cycle for endoreduplication in rodent trophoblast cells.
Mol Biol Cell
9
:
795
807

305

Tan
J
,
Raja
S
,
Davis
MK
,
Tawfik
O
,
Dey
SK
,
Das
SK
2002
Evidence for coordinated interaction of cyclin D3 with p21 and cdk6 in directing the development of uterine stromal cell decidualization and polyploidy during implantation.
Mech Dev
111
:
99
113

306

Tan
Y
,
Li
M
,
Cox
S
,
Davis
MK
,
Tawfik
O
,
Paria
BC
,
Das
SK
2004
HB-EGF directs stromal cell polyploidy and decidualization via cyclin D3 during implantation.
Dev Biol
265
:
181
195

307

Aplin
JD
,
Charlton
AK
,
Ayad
S
1988
An immunohistochemical study of human endometrial extracellular matrix during the menstrual cycle and first trimester of pregnancy.
Cell Tissue Res
253
:
231
240

308

Das
SK
,
Yano
S
,
Wang
J
,
Edwards
DR
,
Nagase
H
,
Dey
SK
1997
Expression of matrix metalloproteinases and tissue inhibitors of metalloproteinases in the mouse uterus during the peri-implantation period.
Dev Genet
21
:
44
54

309

Vu
TH
,
Werb
Z
2000
Matrix metalloproteinases: effectors of development and normal physiology.
Genes Dev
14
:
2123
2133

310

Folkman
J
1995
Angiogenesis in cancer, vascular, rheumatoid and other disease.
Nat Med
1
:
27
31

311

Halder
JB
,
Zhao
X
,
Soker
S
,
Paria
BC
,
Klagsbrun
M
,
Das
SK
,
Dey
SK
2000
Differential expression of VEGF isoforms and VEGF(164)-specific receptor neuropilin-1 in the mouse uterus suggests a role for VEGF(164) in vascular permeability and angiogenesis during implantation.
Genesis
26
:
213
224

312

Hyder
SM
,
Stancel
GM
1999
Regulation of angiogenic growth factors in the female reproductive tract by estrogens and progestins.
Mol Endocrinol
13
:
806
811

313

Chakraborty
I
,
Das
SK
,
Dey
SK
1995
Differential expression of vascular endothelial growth factor and its receptor mRNAs in the mouse uterus around the time of implantation.
J Endocrinol
147
:
339
352

314

Ferrara
N
,
Davis
S
1997
The biology of vascular endothelial growth factor.
Endocr Rev
18
:
4
25

315

Carmeliet
P
,
Ferreira
V
,
Breier
G
,
Pollefeyt
S
,
Kieckens
L
,
Gertsenstein
M
,
Fahrig
M
,
Vandenhoeck
A
,
Harpal
K
,
Eberhardt
C
,
Declercq
C
,
Pawling
J
,
Moons
L
,
Collen
D
,
Risau
W
,
Nagy
A
1996
Abnormal blood vessel development and lethality in embryos lacking a single VEGF allele.
Nature
380
:
435
439

316

Ferrara
N
,
Carver
M
,
Chen
H
,
Dowd
M
,
Lu
L
,
Shea
KS
,
Powell
B
,
Hillan
KJ
,
Moore
MW
1996
Heterozygous embryonic lethality induced by targeted inactivation of the VEGF gene.
Nature
380
:
439
442

317

Tischer
E
,
Mitchell
R
,
Hartman
T
,
Silva
M
,
Gospodarowicz
D
,
Fiddes
JC
,
Abraham
JA
1991
The human gene for vascular endothelial growth factor. Multiple protein forms are encoded through alternative exon splicing.
J Biol Chem
266
:
11947
11954

318

Peters
KG
,
De
V
,
Williams
LT
1993
Vascular endothelial growth factor receptor expression during embryogenesis and tissue repair suggests a role in endothelial differentiation and blood vessel growth.
Proc Natl Acad Sci USA
90
:
8915
8919

319

Shibuya
M
,
Yamaguchi
S
,
Yamane
A
,
Ikeda
T
,
Tojo
A
,
Matsushime
H
,
Sato
M
1990
Nucleotide sequence and expression of a novel human receptor-type tyrosine kinase gene (flt) closely related to the fms family.
Oncogene
5
:
519
524

320

Millauer
B
,
Wizigmann
V
,
Schnurch
H
,
Martinez
R
,
Moller
NP
,
Risau
W
,
Ullrich
A
1993
High affinity VEGF binding and developmental expression suggest Flk-1 as a major regulator of vasculogenesis and angiogenesis.
Cell
72
:
835
846

321

Quinn
TP
,
Peters
KG
,
De
V
,
Ferrara
N
,
Williams
LT
1993
Fetal liver kinase 1 is a receptor for vascular endothelial growth factor and is selectively expressed in vascular endothelium.
Proc Natl Acad Sci USA
90
:
7533
7537

322

Fong
GH
,
Rossant
J
,
Gertsenstein
M
,
Breitman
ML
1995
Role of the Flt-1 receptor tyrosine kinase in regulating the assembly of vascular endothelium.
Nature
376
:
66
70

323

Waltenberger
J
,
Claesson
W
,
Siegbahn
A
,
Shibuya
M
,
Heldin
CH
1994
Different signal transduction properties of KDR and Flt1, two receptors for vascular endothelial growth factor.
J Biol Chem
269
:
26988
26995

324

Yoshida
A
,
Anand
A
,
Zetter
BR
1996
Differential endothelial migration and proliferation to basic fibroblast growth factor and vascular endothelial growth factor.
Growth Factors
13
:
57
64

325

Shalaby
F
,
Rossant
J
,
Yamaguchi
TP
,
Gertsenstein
M
,
Wu
XF
,
Breitman
ML
,
Schuh
AC
1995
Failure of blood-island formation and vasculogenesis in Flk-1-deficient mice.
Nature
376
:
62
66

326

Kawakami
A
,
Kitsukawa
T
,
Takagi
S
,
Fujisawa
H
1996
Developmentally regulated expression of a cell surface protein, neuropilin, in the mouse nervous system.
J Neurobiol
29
:
1
17

327

Takagi
S
,
Kasuya
Y
,
Shimizu
M
,
Matsuura
T
,
Tsuboi
M
,
Kawakami
A
,
Fujisawa
H
1995
Expression of a cell adhesion molecule, neuropilin, in the developing chick nervous system.
Dev Biol
170
:
207
222

328

He
Z
,
Tessier
L
1997
Neuropilin is a receptor for the axonal chemorepellent Semaphorin III.
Cell
90
:
739
751

329

Kolodkin
AL
,
Levengood
DV
,
Rowe
EG
,
Tai
YT
,
Giger
RJ
,
Ginty
DD
1997
Neuropilin is a semaphorin III receptor.
Cell
90
:
753
762

330

Neufeld
G
,
Tessler
S
,
Gitay
G
,
Cohen
T
,
Levi
BZ
1994
Vascular endothelial growth factor and its receptors.
Prog Growth Factor Res
5
:
89
97

331

Soker
S
,
Takashima
S
,
Miao
HQ
,
Neufeld
G
,
Klagsbrun
M
1998
Neuropilin-1 is expressed by endothelial and tumor cells as an isoform-specific receptor for vascular endothelial growth factor.
Cell
92
:
735
745

332

Kawasaki
T
,
Kitsukawa
T
,
Bekku
Y
,
Matsuda
Y
,
Sanbo
M
,
Yagi
T
,
Fujisawa
H
1999
A requirement for neuropilin-1 in embryonic vessel formation.
Development
126
:
4895
4902

333

Kitsukawa
T
,
Shimono
A
,
Kawakami
A
,
Kondoh
H
,
Fujisawa
H
1995
Overexpression of a membrane protein, neuropilin, in chimeric mice causes anomalies in the cardiovascular system, nervous system and limbs.
Development
121
:
4309
4318

334

Hyder
SM
,
Nawaz
Z
,
Chiappetta
C
,
Stancel
GM
2000
Identification of functional estrogen response elements in the gene coding for the potent angiogenic factor vascular endothelial growth factor.
Cancer Res
60
:
3183
3190

335

Ma
W
,
Tan
J
,
Matsumoto
H
,
Robert
B
,
Abrahamson
DR
,
Das
SK
,
Dey
SK
2001
Adult tissue angiogenesis: evidence for negative regulation by estrogen in the uterus.
Mol Endocrinol
15
:
1983
1992

336

Maisonpierre
PC
,
Suri
C
,
Jones
PF
,
Bartunkova
S
,
Wiegand
SJ
,
Radziejewski
C
,
Compton
D
,
McClain
J
,
Aldrich
TH
,
Papadopoulos
N
,
Daly
TJ
,
Davis
S
,
Sato
TN
,
Yancopoulos
GD
1997
Angiopoietin-2, a natural antagonist for Tie2 that disrupts in vivo angiogenesis.
Science
277
:
55
60

337

Suri
C
,
Jones
PF
,
Patan
S
,
Bartunkova
S
,
Maisonpierre
PC
,
Davis
S
,
Sato
TN
,
Yancopoulos
GD
1996
Requisite role of angiopoietin-1, a ligand for the TIE2 receptor, during embryonic angiogenesis.
Cell
87
:
1171
1180

338

Sato
TN
,
Tozawa
Y
,
Deutsch
U
,
Wolburg
B
,
Fujiwara
Y
,
Gendron
M
,
Gridley
T
,
Wolburg
H
,
Risau
W
,
Qin
Y
1995
Distinct roles of the receptor tyrosine kinases Tie-1 and Tie-2 in blood vessel formation.
Nature
376
:
70
74

339

Thurston
G
,
Suri
C
,
Smith
K
,
McClain
J
,
Sato
TN
,
Yancopoulos
GD
,
McDonald
DM
1999
Leakage-resistant blood vessels in mice transgenically overexpressing angiopoietin-1.
Science
286
:
2511
2514

340

Davis
S
,
Yancopoulos
GD
1999
The angiopoietins: Yin and Yang in angiogenesis.
Curr Top Microbiol Immunol
237
:
173
185

341

Valenzuela
DM
,
Griffiths
JA
,
Rojas
J
,
Aldrich
TH
,
Jones
PF
,
Zhou
H
,
McClain
J
,
Copeland
NG
,
Gilbert
DJ
,
Jenkins
NA
,
Huang
T
,
Papadopoulos
N
,
Maisonpierre
PC
,
Davis
S
,
Yancopoulos
GD
1999
Angiopoietins 3 and 4: diverging gene counterparts in mice and humans.
Proc Natl Acad Sci USA
96
:
1904
1909

342

Matsumoto
H
,
Ma
WG
,
Daikoku
T
,
Zhao
X
,
Paria
BC
,
Das
SK
,
Trzaskos
JM
,
Dey
SK
2002
Cyclooxygenase-2 differentially directs uterine angiogenesis during implantation in mice.
J Biol Chem
277
:
29260
29267

343

Gale
NW
,
Thurston
G
,
Hackett
SF
,
Renard
R
,
Wang
Q
,
McClain
J
,
Martin
C
,
Witte
C
,
Witte
MH
,
Jackson
D
,
Suri
C
,
Campochiaro
PA
,
Wiegand
SJ
,
Yancopoulos
GD
2002
Angiopoietin-2 is required for postnatal angiogenesis and lymphatic patterning, and only the latter role is rescued by angiopoietin-1.
Dev Cell
3
:
411
423

344

Rossant
J
,
Howard
L
2002
Signaling pathways in vascular development.
Annu Rev Cell Dev Biol
18
:
541
573

345

Daikoku
T
,
Matsumoto
H
,
Gupta
RA
,
Das
SK
,
Gassmann
M
,
DuBois
RN
,
Dey
SK
2003
Expression of hypoxia-inducible factors in the peri-implantation mouse uterus is regulated in a cell-specific and ovarian steroid hormone-dependent manner. Evidence for differential function of HIFs during early pregnancy.
J Biol Chem
278
:
7683
7691

346

Matsuda
LA
,
Lolait
SJ
,
Brownstein
MJ
,
Young
AC
,
Bonner
TI
1990
Structure of a cannabinoid receptor and functional expression of the cloned cDNA.
Nature
346
:
561
564

347

Munro
S
,
Thomas
KL
,
Abu
S
1993
Molecular characterization of a peripheral receptor for cannabinoids.
Nature
365
:
61
65

348

Devane
WA
,
Hanus
L
,
Breuer
A
,
Pertwee
RG
,
Stevenson
LA
,
Griffin
G
,
Gibson
D
,
Mandelbaum
A
,
Etinger
A
,
Mechoulam
R
1992
Isolation and structure of a brain constituent that binds to the cannabinoid receptor.
Science
258
:
1946
1949

349

Felder
CC
,
Briley
EM
,
Axelrod
J
,
Simpson
JT
,
Mackie
K
,
Devane
WA
1993
Anandamide, an endogenous cannabimimetic eicosanoid, binds to the cloned human cannabinoid receptor and stimulates receptor-mediated signal transduction.
Proc Natl Acad Sci USA
90
:
7656
7660

350

Schmid
PC
,
Paria
BC
,
Krebsbach
RJ
,
Schmid
HH
,
Dey
SK
1997
Changes in anandamide levels in mouse uterus are associated with uterine receptivity for embryo implantation.
Proc Natl Acad Sci USA
94
:
4188
4192

351

Paria
BC
,
Das
SK
,
Dey
SK
1995
The preimplantation mouse embryo is a target for cannabinoid ligand-receptor signaling.
Proc Natl Acad Sci USA
92
:
9460
9464

352

Yang
ZM
,
Paria
BC
,
Dey
SK
1996
Activation of brain-type cannabinoid receptors interferes with preimplantation mouse embryo development.
Biol Reprod
55
:
756
761

353

Wang
J
,
Paria
BC
,
Dey
SK
,
Armant
DR
1999
Stage-specific excitation of cannabinoid receptor exhibits differential effects on mouse embryonic development.
Biol Reprod
60
:
839
844

354

Wang
H
,
Matsumoto
H
,
Guo
Y
,
Paria
BC
,
Roberts
RL
,
Dey
SK
2003
Differential G protein-coupled cannabinoid receptor signaling by anandamide directs blastocyst activation for implantation.
Proc Natl Acad Sci USA
100
:
14914
14919

355

Maccarrone
M
,
Valensise
H
,
Bari
M
,
Lazzarin
N
,
Romanini
C
,
Finazzi-Agro
A
2000
Relation between decreased anandamide hydrolase concentrations in human lymphocytes and miscarriage.
Lancet
355
:
1326
1329

356

Maccarrone
M
,
Bisogno
T
,
Valensise
H
,
Lazzarin
N
,
Fezza
F
,
Manna
C
,
Di
M
,
V, Finazzi-Agro
A
2002
Low fatty acid amide hydrolase and high anandamide levels are associated with failure to achieve an ongoing pregnancy after IVF and embryo transfer.
Mol Hum Reprod
8
:
188
195

357

Ingham
PW
,
McMahon
AP
2001
Hedgehog signaling in animal development: paradigms and principles.
Genes Dev
15
:
3059
3087

358

Johnson
RL
,
Scott
MP
1998
New players and puzzles in the hedgehog signaling pathway.
Curr Opin Genet Dev
8
:
450
456

359

McMahon
AP
2000
More surprises in the hedgehog signaling pathway.
Cell
100
:
185
188

360

Matsumoto
H
,
Zhao
X
,
Das
SK
,
Hogan
BL
,
Dey
SK
2002
Indian hedgehog as a progesterone-responsive factor mediating epithelial-mesenchymal interactions in the mouse uterus.
Dev Biol
245
:
280
290

361

Takamoto
N
,
Zhao
B
,
Tsai
SY
,
DeMayo
FJ
2002
Identification of Indian hedgehog as a progesterone-responsive gene in the murine uterus.
Mol Endocrinol
16
:
2338
2348

362

Okulicz
WC
,
Ace
CI
1999
Progesterone-regulated gene expression in the primate endometrium.
Semin Reprod Endocrinol
17
:
241
255

363

Higuchi
T
,
Kanzaki
H
,
Nakayama
H
,
Fujimoto
M
,
Hatayama
H
,
Kojima
K
,
Iwai
M
,
Mori
T
,
Fujita
J
1995
Induction of tissue inhibitor of metalloproteinase 3 gene expression during in vitro decidualization of human endometrial stromal cells.
Endocrinology
136
:
4973
4981

364

Bany
BM
,
Schultz
GA
2001
Increased expression of a novel heat shock protein transcript in the mouse uterus during decidualization and in response to progesterone.
Biol Reprod
64
:
284
292

365

Simmons
DG
,
Kennedy
TG
2002
Uterine sensitization-associated gene-1: a novel gene induced within the rat endometrium at the time of uterine receptivity/sensitization for the decidual cell reaction.
Biol Reprod
67
:
1638
1645

366

Gruidl
M
,
Buyuksal
A
,
Babaknia
A
,
Fazleabas
AT
,
Sivarajah
S
,
Satyaswaroop
PG
,
Tabibzadeh
S
1997
The progressive rise in the expression of α crystallin B chain in human endometrium is initiated during the implantation window: modulation of gene expression by steroid hormones.
Mol Hum Reprod
3
:
333
342

367

Green
ML
,
Blaeser
LL
,
Simmen
FA
,
Simmen
RC
1996
Molecular cloning of spermidine/spermine N1-acetyltransferase from the periimplantation porcine uterus by messenger ribonucleic acid differential display: temporal and conceptus-modulated gene expression.
Endocrinology
137
:
5447
5455

368

Cowan
BD
,
Hines
RS
,
Brackin
MN
,
Case
ST
1999
Temporal and cell-specific gene expression by human endometrium after coculture with trophoblast.
Am J Obstet Gynecol
180
:
806
814

369

Nie
GY
,
Li
Y
,
Hampton
AL
,
Salamonsen
LA
,
Clements
JA
,
Findlay
JK
2000
Identification of monoclonal nonspecific suppressor factor β (mNSFβ) as one of the genes differentially expressed at implantation sites compared to interimplantation sites in the mouse uterus.
Mol Reprod Dev
55
:
351
363

370

Simmons
DG
,
Kennedy
TG
2000
Induction of glucose-regulated protein 78 in rat uterine glandular epithelium during uterine sensitization for the decidual cell reaction.
Biol Reprod
62
:
1168
1176

371

Martin
J
,
Dominguez
F
,
Avila
S
,
Castrillo
JL
,
Remohi
J
,
Pellicer
A
,
Simon
C
2002
Human endometrial receptivity: gene regulation.
J Reprod Immunol
55
:
131
139

372

Lee
S
,
Lee
SA
,
Shim
C
,
Khang
I
,
Lee
KA
,
Park
YM
,
Kang
BM
,
Kim
K
2003
Identification of estrogen-regulated genes in the mouse uterus using a delayed-implantation model.
Mol Reprod Dev
64
:
405
413

373

Evans
SJ
,
Datson
NA
,
Kabbaj
M
,
Thompson
RC
,
Vreugdenhil
E
,
De
K
,
Watson
SJ
,
Akil
H
2002
Evaluation of Affymetrix Gene Chip sensitivity in rat hippocampal tissue using SAGE analysis. Serial analysis of gene expression.
Eur J Neurosci
16
:
409
413

374

Ishii
M
,
Hashimoto
S
,
Tsutsumi
S
,
Wada
Y
,
Matsushima
K
,
Kodama
T
,
Aburatani
H
2000
Direct comparison of GeneChip and SAGE on the quantitative accuracy in transcript profiling analysis.
Genomics
68
:
136
143

375

Reese
J
,
Das
SK
,
Paria
BC
,
Lim
H
,
Song
H
,
Matsumoto
H
,
Knudtson
KL
,
DuBois
RN
,
Dey
SK
2001
Global gene expression analysis to identify molecular markers of uterine receptivity and embryo implantation.
J Biol Chem
276
:
44137
44145

376

Yoshioka
K
,
Matsuda
F
,
Takakura
K
,
Noda
Y
,
Imakawa
K
,
Sakai
S
2000
Determination of genes involved in the process of implantation: application of GeneChip to scan 6500 genes.
Biochem Biophys Res Commun
272
:
531
538

377

Cheon
YP
,
Li
Q
,
Xu
X
,
DeMayo
FJ
,
Bagchi
IC
,
Bagchi
MK
2002
A genomic approach to identify novel progesterone receptor regulated pathways in the uterus during implantation.
Mol Endocrinol
16
:
2853
2871

378

Andrade
PM
,
Silva
ID
,
Borra
RC
,
de Lima
GR
,
Baracat
EC
2002
Estrogen regulation of uterine genes in vivo detected by complementary DNA array.
Horm Metab Res
34
:
238
244

379

Watanabe
H
,
Suzuki
A
,
Mizutani
T
,
Khono
S
,
Lubahn
DB
,
Handa
H
,
Iguchi
T
2002
Genome-wide analysis of changes in early gene expression induced by oestrogen.
Genes Cells
7
:
497
507

380

Lindberg
MK
,
Weihua
Z
,
Andersson
N
,
Moverare
S
,
Gao
H
,
Vidal
O
,
Erlandsson
M
,
Windahl
S
,
Andersson
G
,
Lubahn
DB
,
Carlsten
H
,
Dahlman
W
,
Gustafsson
JA
,
Ohlsson
C
2002
Estrogen receptor specificity for the effects of estrogen in ovariectomized mice.
J Endocrinol
174
:
167
178

381

Naciff
JM
,
Jump
ML
,
Torontali
SM
,
Carr
GJ
,
Tiesman
JP
,
Overmann
GJ
,
Daston
GP
2002
Gene expression profile induced by 17α-ethynyl estradiol, bisphenol A, and genistein in the developing female reproductive system of the rat.
Toxicol Sci
68
:
184
199

382

Brar
AK
,
Handwerger
S
,
Kessler
CA
,
Aronow
BJ
2001
Gene induction and categorical reprogramming during in vitro human endometrial fibroblast decidualization.
Physiol Genomics
7
:
135
148

383

Popovici
RM
,
Kao
LC
,
Giudice
LC
2000
Discovery of new inducible genes in in vitro decidualized human endometrial stromal cells using microarray technology.
Endocrinology
141
:
3510
3513

384

Kao
LC
,
Tulac
S
,
Lobo
S
,
Imani
B
,
Yang
JP
,
Germeyer
A
,
Osteen
K
,
Taylor
RN
,
Lessey
BA
,
Giudice
LC
2002
Global gene profiling in human endometrium during the window of implantation.
Endocrinology
143
:
2119
2138

385

Carson
DD
,
Lagow
E
,
Thathiah
A
,
Al
S
,
Farach
C
,
Vernon
M
,
Yuan
L
,
Fritz
MA
,
Lessey
B
2002
Changes in gene expression during the early to mid-luteal (receptive phase) transition in human endometrium detected by high-density microarray screening.
Mol Hum Reprod
8
:
871
879

386

Weitlauf
HM
1994
Biology of implantation. In:
Knobil
E
,
Neill
JD
, eds. The physiology of reproduction.
New York
:
Raven Press
;
391
440

387

Yoshinaga
K
1972
Rabbit antiserum to rat deciduoma.
Biol Reprod
6
:
51
57

388

Denari
JH
,
Germino
NI
,
Rosner
JM
1976
Early synthesis of uterine proteins after a decidual stimulus in the pseudopregnant rat.
Biol Reprod
15
:
1
8

389

Bell
SC
,
Reynolds
S
,
Heald
PJ
1977
Uterine protein synthesis during the early stages of pregnancy in the rat.
J Reprod Fertil
49
:
177
181

390

Umapathysivam
K
,
Jones
WR
1978
An investigation of decidual specific proteins in the rat.
Int J Fertil
23
:
138
142

391

Lejeune
B
,
Lecocq
R
,
Lamy
F
,
Leroy
F
1982
Changes in the pattern of endometrial protein synthesis during decidualization in the rat.
J Reprod Fertil
66
:
519
523

392

Mulholland
J
,
Villee
CA
1984
Proteins synthesized by the rat endometrium during early pregnancy.
J Reprod Fertil
72
:
395
400

393

Salamonsen
LA
,
Doughton
BW
,
Findlay
JK
1986
The effects of the preimplantation blastocyst in vivo and in vitro on protein synthesis and secretion by cultured epithelial cells from sheep endometrium.
Endocrinology
119
:
622
628

394

Weitlauf
HM
,
Suda
H
1988
Changes in secreted uterine proteins associated with embryo implantation in the mouse.
J Reprod Fertil
84
:
539
549

395

Lee
RS
,
Wheeler
TT
,
Peterson
AJ
1998
Large-format, two-dimensional polyacrylamide gel electrophoresis of ovine periimplantation uterine luminal fluid proteins: identification of aldose reductase, cytoplasmic actin, and transferrin as conceptus-synthesized proteins.
Biol Reprod
59
:
743
752

396

Banks
RE
,
Dunn
MJ
,
Forbes
MA
,
Stanley
A
,
Pappin
D
,
Naven
T
,
Gough
M
,
Harnden
P
,
Selby
PJ
The potential use of laser capture microdissection to selectively obtain distinct populations of cells for proteomic analysis–preliminary findings.
Electrophoresis
20
:
689
700
.

397

Hoang
VM
,
Foulk
R
,
Clauser
K
,
Burlingame
A
,
Gibson
BW
,
Fisher
SJ
2001
Functional proteomics: examining the effects of hypoxia on the cytotrophoblast protein repertoire.
Biochemistry
40
:
4077
4086

398

Cencic
A
,
Henry
C
,
Lefevre
F
,
Huet
JC
,
Koren
S
,
La
B
2002
The porcine trophoblastic interferon-γ, secreted by a polarized epithelium, has specific structural and biochemical properties.
Eur J Biochem
269
:
2772
2781

399

Bazer
FW
,
Vallet
JL
,
Ashworth
CJ
,
Anthony
RV
,
Roberts
RM
1987
The role of ovine conceptus secretory proteins in the establishment of pregnancy.
Adv Exp Med Biol
230
:
221
235

400

Perry
JS
,
Heap
RB
,
Amoroso
EC
1973
Steroid hormone production by pig blastocysts.
Nature
245
:
45
47

401

Torres
M
,
Stoykova
A
,
Huber
O
,
Chowdhury
K
,
Bonaldo
P
,
Mansouri
A
,
Butz
S
,
Kemler
R
,
Gruss
P
1997
An α-E-catenin gene trap mutation defines its function in preimplantation development.
Proc Natl Acad Sci USA
94
:
901
906

402

Christiansen
W
,
Voland
JR
,
Wu
Y
,
Ngo
K
,
Roland
BL
,
Nguyen
S
,
Peterson
PA
,
Fung
L
2000
Functional loss of ABCA1 in mice causes severe placental malformation, aberrant lipid distribution, and kidney glomerulonephritis as well as high-density lipoprotein cholesterol deficiency.
Am J Pathol
157
:
1017
1029

403

Feldman
B
,
Poueymirou
W
,
Papaioannou
VE
,
DeChiara
TM
,
Goldfarb
M
1995
Requirement of FGF-4 for postimplantation mouse development.
Science
267
:
246
249

404

Xiao
X
,
Zuo
X
,
Davis
AA
,
McMillan
DR
,
Curry
BB
,
Richardson
JA
,
Benjamin
IJ
1999
HSF1 is required for extra-embryonic development, postnatal growth and protection during inflammatory responses in mice.
EMBO J
18
:
5943
5952

405

Christians
E
,
Davis
AA
,
Thomas
SD
,
Benjamin
IJ
2000
Maternal effect of Hsf1 on reproductive success.
Nature
407
:
693
694

406

Tong
ZB
,
Gold
L
,
Pfeifer
KE
,
Dorward
H
,
Lee
E
,
Bondy
CA
,
Dean
J
,
Nelson
LM
2000
Mater, a maternal effect gene required for early embryonic development in mice.
Nat Genet
26
:
267
268

407

Burns
KH
,
Viveiros
MM
,
Ren
Y
,
Wang
P
,
DeMayo
FJ
,
Frail
DE
,
Eppig
JJ
,
Matzuk
MM
2003
Roles of NPM2 in chromatin and nucleolar organization in oocytes and embryos.
Science
300
:
633
636

408

Yuan
L
,
Liu
JG
,
Hoja
MR
,
Wilbertz
J
,
Nordqvist
K
,
Hoog
C
2002
Female germ cell aneuploidy and embryo death in mice lacking the meiosis-specific protein SCP3.
Science
296
:
1115
1118

409

Shindo
T
,
Kurihara
H
,
Kuno
K
,
Yokoyama
H
,
Wada
T
,
Kurihara
Y
,
Imai
T
,
Wang
Y
,
Ogata
M
,
Nishimatsu
H
,
Moriyama
N
,
Oh-hashi
Y
,
Morita
H
,
Ishikawa
T
,
Nagai
R
,
Yazaki
Y
,
Matsushima
K
2000
ADAMTS-1: a metalloproteinase-disintegrin essential for normal growth, fertility, and organ morphology and function.
J Clin Invest
105
:
1345
1352

410

Terasawa
Y
,
Ladha
Z
,
Leonard
SW
,
Morrow
JD
,
Newland
D
,
Sanan
D
,
Packer
L
,
Traber
MG
,
Farese
RV
2000
Increased atherosclerosis in hyperlipidemic mice deficient in α-tocopherol transfer protein and vitamin E.
Proc Natl Acad Sci USA
97
:
13830
13834

411

Kuno
N
,
Kadomatsu
K
,
Fan
QW
,
Hagihara
M
,
Senda
T
,
Mizutani
S
,
Muramatsu
T
1998
Female sterility in mice lacking the basigin gene, which encodes a transmembrane glycoprotein belonging to the immunoglobulin superfamily.
FEBS Lett
425
:
191
194

412

Igakura
T
,
Kadomatsu
K
,
Kaname
T
,
Muramatsu
H
,
Fan
QW
,
Miyauchi
T
,
Toyama
Y
,
Kuno
N
,
Yuasa
S
,
Takahashi
M
,
Senda
T
,
Taguchi
O
,
Yamamura
K
,
Arimura
K
,
Muramatsu
T
1998
A null mutation in basigin, an immunoglobulin superfamily member, indicates its important roles in peri-implantation development and spermatogenesis.
Dev Biol
194
:
152
165

413

Fowler
KJ
,
Hudson
DF
,
Salamonsen
LA
,
Edmondson
SR
,
Earle
E
,
Sibson
MC
,
Choo
KH
2000
Uterine dysfunction and genetic modifiers in centromere protein B-deficient mice.
Genome Res
10
:
30
41

414

Cohen
PE
,
Zhu
L
,
Pollard
JW
1997
Absence of colony stimulating factor-1 in osteopetrotic (csfmop/csfmop) mice disrupts estrous cycles and ovulation.
Biol Reprod
56
:
110
118

415

Hess
RA
,
Bunick
D
,
Lee
KH
,
Bahr
J
,
Taylor
JA
,
Korach
KS
,
Lubahn
DB
1997
A role for oestrogens in the male reproductive system.
Nature
390
:
509
512

416

Ogawa
S
,
Eng
V
,
Taylor
J
,
Lubahn
DB
,
Korach
KS
,
Pfaff
DW
1998
Roles of estrogen receptor-α gene expression in reproductionrelated behaviors in female mice.
Endocrinology
139
:
5070
5081

417

Ogawa
S
,
Lubahn
DB
,
Korach
KS
,
Pfaff
DW
1997
Behavioral effects of estrogen receptor gene disruption in male mice.
Proc Natl Acad Sci USA
94
:
1476
1481

418

Satokata
I
,
Benson
G
,
Maas
R
1995
Sexually dimorphic sterility phenotypes in Hoxa10-deficient mice.
Nature
374
:
460
463

419

Panda
DK
,
Miao
D
,
Tremblay
ML
,
Sirois
J
,
Farookhi
R
,
Hendy
GN
,
Goltzman
D
2001
Targeted ablation of the 25-hydroxyvitamin D 1α-hydroxylase enzyme: evidence for skeletal, reproductive, and immune dysfunction.
Proc Natl Acad Sci USA
98
:
7498
7503

420

Baker
J
,
Hardy
MP
,
Zhou
J
,
Bondy
C
,
Lupu
F
,
Bellve
AR
,
Efstratiadis
A
1996
Effects of an Igf1 gene null mutation on mouse reproduction.
Mol Endocrinol
10
:
903
918

421

Smith
CL
,
DeVera
DG
,
Lamb
DJ
,
Nawaz
Z
,
Jiang
YH
,
Beaudet
AL
,
O’Malley
BW
2002
Genetic ablation of the steroid receptor coactivator-ubiquitin ligase, E6-AP, results in tissue-selective steroid hormone resistance and defects in reproduction.
Mol Cell Biol
22
:
525
535

422

Parr
BA
,
McMahon
AP
1998
Sexually dimorphic development of the mammalian reproductive tract requires Wnt-7a.
Nature
395
:
707
710