HHS Public Access Author manuscript Author Manuscript

Reproduction. Author manuscript; available in PMC 2017 November 01. Published in final edited form as: Reproduction. 2016 November ; 152(5): R179–R189. doi:10.1530/REP-16-0325.

The Evolution of the Placenta R Michael Roberts1,2, Jonathan A Green2, and Laura C Schulz3 1C.S.

Bond Life Sciences Center, University of Missouri, Columbia, Missouri

2Division

of Animal Sciences, University of Missouri, Columbia, Missouri

3Department

Author Manuscript

of Obstetrics, Gynecology and Women’s Health, University of Missouri School of Medicine, Columbia, Missouri

Abstract

Author Manuscript

The still apt definition of a placenta is that coined by Mossman, namely apposition or fusion of the fetal membranes to the uterine mucosa for physiological exchange. As such it is a specialized organ whose purpose is to provide continuing support to the developing young. By this definition, placentas have evolved within every vertebrate class other than birds. They have evolved on multiple occasions, often within quite narrow taxonomic groups. As the placenta and the maternal system associate more intimately, such that the conceptus relies extensively on maternal support, the relationship leads to increased conflict that drives adaptive changes on both sides. The story of vertebrate placentation, therefore, is one of convergent evolution at both the macro- and molecular levels. In this short review, we first describe the emergence of placental-like structures in nonmammalian vertebrates and then transition to mammals themselves. We close the review by discussing mechanisms that might have favored diversity and hence evolution of the morphology and physiology of the placentas of eutherian mammals.

Introduction

Author Manuscript

Viviparity has evolved independently and seemingly on multiple occasions across a diverse array of animal groups, including invertebrates(Kalinka 2015). It is a phenomenon whereby developing embryos are retained within the reproductive tract, leading to release of live offspring as an alternative to the more fecund egg laying or spawning. One consequence of viviparity is that the retained, fertilized egg must either survive off its own reserves, usually yolk, or obtain some or part of these resources from the mother. The latter situation, of necessity, is expected to lead to increased conflict over how provisions are partitioned between the supplier and the recipient, thus potentially sparking a genetic arms race. In turn, such conflict is expected to drive adaptive changes that lead to a more intimate and possibly more felicitous relationship between offspring and the reproductive tract of the mother that favors the transmission of both maternal and paternal genes to the next generation, despite a loss of overall fecundity.

Correspondence should be addressed to R Michael Roberts; [email protected]. Declaration of interest The authors declare that there is no conflict of interest that could be perceived as prejudicing the impartiality of the research reported.

Roberts et al.

Page 2

Author Manuscript

The transition from oviparity to viviparity and the subsequent emergence of placentation within some vertebrate taxa clearly required major changes in the morphology and physiology of the reproductive tract and has its origins well before the advent of mammals (Blackburn 2015, Van Dyke, et al. 2014). A placenta, as defined originally by Mossman, is the “apposition or fusion of the fetal membranes to the uterine mucosa for physiological exchange” (Burton and Jauniaux 2015). It is a specialized organ whose purpose is to provide continuing support to the developing young, through the provision of water, nutrients, and gasses, and to regulate maternal-fetal interactions often through hormone production. By this definition, placentas have evolved within every vertebrate class other than birds. Although placentation arose once in the common ancestor of mammals, it has arisen independently multiple times within other classes, and even families. The story of vertebrate placentation, therefore, is one of convergent evolution at both the macro- and molecular levels. In this short review of placental evolution, we first describe the emergence of placental-like structures in non-mammalian vertebrates and then transition to mammals themselves. We close the review by discussing mechanisms that might have favored diversity and hence evolution of the morphology and physiology of the placentas of eutherian mammals. Of necessity, many important references cannot be cited in a short review of this kind. Instead we have attempted to direct the reader to scholarly articles that do list the primary source material.

Author Manuscript

Non-Mammalian Vertebrates Cartilagenous Fish

Author Manuscript

Viviparity is the most common mode of reproduction in elasmobranchs, but there is a wide range in the degree of maternal provisioning of the embryo after ovulation (Wourms and Lombardi 1992), with embryos of some species depending entirely on the yolk sac for all of pregnancy, e.g. the spiny dogfish, which at birth weighs 40% less than the yolk-filled fertilized egg from which it develops. The tiger shark, by contrast, solves such limiting provisioning by the dominant embryo eating all of its littermates and any unfertilized eggs before birth (Korsgaard and Weber 1989).

Author Manuscript

In the elasmobranchs, two placental types are observed. Among stingrays, fingerlike projections of the uterine wall, termed trophonemata, provide histotrophic nutrition to the developing embryo (Hamlett, et al. 1993), while the epithelium overlying uterine blood vessels thins, lessening the barrier to exchange. The mature embryo of the cownose ray, for example, weighs 3000 times as much as the egg as a result of reliance on the mother rather than the egg yolk (Hamlett, et al. 1993). A different placental type, the yolk sac placenta, has evolved multiple times in the ground sharks(Wourms and Lombardi 1992). After the yolk has been consumed, the yolk sac becomes modified into an umbilical cord region and a placental region (Hamlett, et al. 1993), which becomes closely apposed to the epithelium of a highly vascularized oviduct. Teleosts In Poeciliopsis fish, eggs are retained within the follicle after fertilization and throughout embryonic development (Wourms and Lombardi 1992). Most species are lecithotrophic,

Reproduction. Author manuscript; available in PMC 2017 November 01.

Roberts et al.

Page 3

Author Manuscript

relying on yolk, but in others, a highly folded, follicular epithelium forms a “follicular placenta” with the embryonic pericardial sac. This epithelial layer is lined with microvilli, while the surrounding tissue is highly vascularized, presumably to facilitate maternal-fetal exchange (Kwan, et al. 2015). This type of placental structure has arisen several times during Poeciliopsis evolution (Kwan, et al. 2015) and has been subsequently lost and regained many times, making it a model for the evolution of placentation (Pollux, et al. 2014).

Author Manuscript

Among Syngnathid fish (pipefish and seahorses) the males and not the females become pregnant, after the females transfer the fertilized eggs to a brood pouch on the ventral side of either the trunk (Gastrophori) or the tail (Urophori) of the males’ body. The degree of contact and exchange between developing embryos and the father range from minimal to situations that display all the defining features of placentation (Wilson, et al. 2001). For example, in the straight-nosed pipefish Nerophis ophidion, eggs stick to the epithelium of the father’s pouch, which is open to the sea environment (Carcupino, et al. 2002), while in the potbellied seahorse Hippocampus abdominalis, the pouch is sealed and its epithelium highly folded, thinned, and vascularized, (Carcupino, et al. 2002), thereby allowing exchange of nutrients. Amphibians

Author Manuscript

Viviparity has arisen multiple times in amphibians, accompanied in some cases with the evolution of placentas. In frogs, there are examples in which the tadpoles develop in the father’s mouth, in the mother’s stomach, or on the skin of the back (Wake 1993). In the marsupial frogs, development occurs inside a specialized maternal pouch on the back of the animal. Following ovulation, secretory cells of the vascularized pouch enlarge. The pouch epithelium is closely apposed to specialized fetal gills, allowing exchange between maternal and fetal circulations (Savage 2002). This, as A.M. Carter wrote, “would satisfy most people’s definition of a placenta” (Carter 2016). Arguably, however, a structure found outside the reproductive tract, especially following external fertilization as occurs in these frogs and the sea horses described earlier, does not fit the typical definition of a placenta. Nor have any of the “gill placentas” of these endangered frog species yet been shown to be the primary source of fetal nutrient support. Reptiles

Author Manuscript

Viviparity occurs in more than 25 families of lizards and snakes (squamate reptiles). Although estimated by some to have arisen independently more than one hundred times, it is conceivable that there was a single origin of viviparity and multiple subsequent reversions to oviparity (Blackburn 2015, Pyron and Burbrink 2014, Van Dyke, et al. 2014). In many species, the embryo, relies on egg contents for nutrition, but, in others, a range of adaptations of the female reproductive tract provides a means to exchange gases and nutrients with the conceptus (Stewart 2015). In squamates the placenta is chorioallantoic, but unlike in mammals, does not develop from an early arising, extraembryonic, trophoblast layer. Most squamate placentas demonstrate a simple interdigitation of the chorioallantois with the uterine epithelium (Stewart 2015), but in some skinks the interface is more intimate (Fig. 1). In the case of the African skink T. ivensi, there is even a degree of invasive

Reproduction. Author manuscript; available in PMC 2017 November 01.

Roberts et al.

Page 4

Author Manuscript

implantation, with direct contact between chorionic projections and maternal capillary endothelium (Blackburn and Flemming 2012). The chorioallantoic placenta of viviparous lizards can be highly efficient in facilitating exchange of nutrients and gases. Offspring of the Mabuya lizard weigh 500 times more at birth than the egg at fertilization, indicating a nearly complete reliance for growth on maternally supplied materials (Thompson and Speake 2002). There is also evidence for expression of active transporters at the sites of maternal-fetal apposition in some species (Murphy, et al. 2011) and for placental production of steroid hormones (Painter and Moore 2005). Overall, where placentation occurs in squamates, it bears a superficial resemblance to the epitheliochorial placentation encountered in some eutherian mammals and carries out many of the same functions, but its formation does not involve the formation of a trophoblast lineage, a feature unique to placental mammals.

Author Manuscript

Mammals

Author Manuscript

All mammals except the egg-laying platypus and the five species of echidnas, the only surviving monotremes, rely on a placenta for their reproduction. Indeed, the first lineage decision made during embryonic development of Mammalia is the segregation of cells destined to become the external tissue layer of the placenta. This lineage, usually called trophoblast (or, because it forms the exterior of the conceptus, trophectoderm) diverges early from the pluripotent lineage that advances to form the inner cell mass, the cells that give rise to the hypoblast and epiblast, is unique to mammals. The term trophoblast was first coined by Hubrecht (Hubrecht 1904). Its roots are “tropho” meaning nourishment, and “blast” for embryonic. Whether embryo-derived placental cells of other animals, which, unlike mammals, do not form an initiating trophectoderm, should be termed “trophoblast” is therefore questionable, despite many shared functions. When and how trophoblast emerged in the presumed metatherian ancestors of the Class Mammalia is mysterious and unlikely ever to be revealed from the fossil record. Marsupials

Author Manuscript

Although eutherians are often referred to as “placental mammals”, marsupials do possess placentas. However, the placentas are short-lived and marsupials, birth relatively underdeveloped young. In the marsupial blastocyst, blastomeres adhere to the inner surface of the zona pellucida, resulting first in a cup-shaped blastocyst, and then a hollow sphere, with no discernible inner cell mass (Selwood and Johnson 2006). Both the embryonic and extraembryonic lineages emerge from this single cell layer, with trophoblast materializing from the pole of the blastocyst opposite the position of the future embryo (Familari, et al. 2016, Frankenberg, et al. 2013, Morrison, et al. 2013). There are also indications that early lineage specification may be controlled by a similar set of transcription factors as in eutherians (Familari, et al. 2016, Frankenberg, et al. 2013, Morrison, et al. 2013). The trophoblast cells originating from the marsupial blastocyst combine with yolk sac endoderm and extra-embryonic mesoderm to form a yolk sac placenta, which can be bi- or trilaminar (Renfree 2010). There are a species of koalas, wombats, and bandicoots, however, in which a chorioallantoic placenta forms, but it appears to be supplementary to the main,

Reproduction. Author manuscript; available in PMC 2017 November 01.

Roberts et al.

Page 5

Author Manuscript

yolk sac placenta(Freyer, et al. 2003). Although the placentas of marsupials are generally regarded as non-invasive and of the epitheliochorial type (see next section), an area of syncytium forms in the yolk sac placenta of the gray short-tailed opossum and possibly related species. This tissue makes direct contact with maternal blood vessels (Zeller and Freyer 2001). Additionally, trophoblast syncytialization (syncytial trophoblast; STB), where adjacent cells fuse to produce cells with more than a single nucleus, is accompanied by expression of an endogenous, retrovirus-derived, protein, analogous to what occurs in many eutherians (Cornelis, et al. 2014). In another similarity to the eutherians, the placenta of the tammar wallaby secretes a complement of hormones that include IGF2 and relaxin, which, in that species, may be responsible for the phenomenon of maternal recognition of pregnancy (Renfree 2010). Eutherians

Author Manuscript Author Manuscript Author Manuscript

Origin of trophoblast—Eutherians consist of 5,000 or so species in roughly 20 phylogenetic orders. Most of our knowledge about their early development and the events that lead to the initial specification of trophoblast, has been obtained from only a handful of these species, especially the mouse. In the mouse (Mus musculus), the first visible differentiation event, called compaction, occurs between the 8- and 16-cell stage of embryonic development when the blastomeres polarize to form extended contact zones with their neighbors and an outward-facing apical surface (Posfai, et al. 2014). Further symmetrical and asymmetrical divisions follow, so that by the 16-cell morula stage there exists a population of polarized cells on the cell surface, which becomes trophectoderm, and another population positioned inside, which is the precursor of the inner cell mass (ICM). By the time of blastocyst formation, when the mouse conceptus comprises about 32 cells, about two-thirds of which are trophectoderm, the precursors of extraembryonic endoderm (hypoblast) have begun to segregate towards the base of the ICM. From this location they migrate outwards to line the trophectoderm, forming the yolk sac cavity. Later, extraembryonic mesoderm joins with the trophectoderm to form what at this stage is a bilayered chorion. Although there are clear morphological similarities in the steps leading to blastocyst formation in rodents, humans, and domestic mammalian species, there are differences in how events are timed and subsequently play out. In many species,, for example, mesoderm differentiates to vascularize the yolk sac, forming a yolk sac placenta that is eventually replaced by the chorioallantoic placenta. Such a placenta does not form in humans, but persists to term in rodents and lagomorphs. Similarly, compared to the mouse, conceptuses from many species undergo at least one further round of cell division before the blastocoel cavity becomes visible. For details on how placentas subsequently develop and begin to diverge in morphology, the reader is directed to the still germane reviews by Amoroso (Amoroso 1952) and Renfree (Renfree 1982). Range of placental morphologies—Despite the spectacular diversity encountered in placental structure at both the gross and microanatomical levels (Enders and Carter 2004), there have been attempts over the years to classify placentas and (somewhat unsuccessfully) to relate the outcomes to mammalian phylogeny. For example, placentas come in a range of shapes and sizes and have been placed into groups based on the general gross morphologies of the sites where the chorion attaches to the endometrium (Fig. 2). They have also been

Reproduction. Author manuscript; available in PMC 2017 November 01.

Roberts et al.

Page 6

Author Manuscript Author Manuscript

classified according to how many cell layers separate the maternal and fetal circulations (Fig. 2). A simplified version of this system recognizes three main placental types (Renfree 1982): epitheliochorial, in which there is no erosion of the uterine epithelium; endotheliochorial, where the invading trophoblast reaches but does not penetrate maternal capillaries within the endometrium; and haemochorial, where the trophoblast surface is in direct contact with maternal blood (Fig. 3). A fourth type, synepitheliochorial, has been recently used to describe placentas of ruminants (Wooding 1992). Here, specialized trophoblast binucleated cells fuse with uterine epithelial cells to form trinucleated cells, a syncytium, or both. The placentas of ruminants, with the exception of the phylogenetically more primitive tragulids, are also characterized as cotyledonary (Fig. 2). Cotyledons are the primary sites of attachment of the chorion to the endometrium and comprise tufts of villous trophoblast that interdigitate and interlock with complementary structures on the endometrium, called caruncles. It is within these sites, often known as placentomes, that binucleated cells, placental and maternal blood capillaries, and placental interchange are most concentrated. This kind of placenta is, therefore, a derived form, secondary to the more superficial, diffuse epitheliochorial type typical of most other artiodactyls.

Author Manuscript

Superficial versus invasive placentation—Intuitively, it might be inferred that the least invasive types of placenta, the kind where multiple cell layers separate the two blood supplies (Fig. 2B), would be the most inefficient and also the most primitive of all placental types because there are more apparent physical barriers to limit movement of nutrients and dissolved gases. Also such placentas have a resemblance to placentas described earlier for non-mammalian species, such as skinks (Fig. 3) However, this is clearly not the case (Wildman, et al. 2006). Compare, for example, the relative maturity and independence of a newborn horse, pig or whale, supported during their gestations by diffuse epitheliochorial placentas, to a hapless pup born to a mouse or a baby born to a human, species where the placenta is hemochorial and discoid. Nor is the more superficial form of placenta evolutionarily more ancient than the kinds that are invasive and make direct contact with maternal blood. On the contrary, careful analyses, in which placental features were mapped to mammalian phylogenies, have clearly dispelled this myth and demonstrated that the epitheliochorial placenta is not only a derived form that has evolved from a more invasive placental type, but has arisen independently in three distinct mammalian lineages, including primates (Wildman, et al. 2006) (Fig. 4). Moreover, this and similar trees based on other features of the placenta strongly imply that the placenta of the last common eutherian ancestor was discoid, either hemochorial or endotheliochorial (Mess and Carter 2007), and possessed a labyrinth-type type of placental interdigitation.

Author Manuscript

It should be stressed that the perceived disadvantages of extended diffusion distances for dissolved gases and solutes encountered with a superficial placenta, like that of the pig, are largely illusory. As pregnancy proceeds, the surfaces of the uterine and trophoblast epithelial layers become interlocked (Fig. 5B). Capillaries on the maternal and fetal sides come within a few microns of each other (Fig. 5A), often squeezing down to the tight junctions between the trophoblast cells on the one side and uterine epithelial cells on the other, probably allowing efficient exchange of small molecules (Friess, et al. 1980). Movement of macromolecules that carry essential components such as metals and vitamins, is more

Reproduction. Author manuscript; available in PMC 2017 November 01.

Roberts et al.

Page 7

Author Manuscript Author Manuscript

problematic, but there an alternative strategy has been employed, namely the delivery of these components, not in the blood but in uterine secretions, a process called histotrophic nutrition (Fig. 5C) (Renegar, et al. 1982). The uterine glands of the pig, for example, release uteroferrin, a bi-iron containing acid phosphatase to supply iron. Uteroferrin and other proteins with a similar provisioning function are taken up by specialized regions of endocytic trophoblast cells congregated in cup-like structures called areolae, which develop opposite the mouths of each uterine gland (Fig. 5C). From there, the maternal factors are transported via the fetal bloodstream to the liver and other organs for processes such as hematopoiesis (Renegar, et al. 1982). Among the likely advantages of a non-invasive placenta over the more primitive hemochorial and endotheliochorial types include reduced exposure to the potential threats of the maternal immune system, less damage to the uterus associated with birth and elimination of the placenta, and minimizing the transmission of fetal cells into the maternal blood circulation and vice versa. Finally, superficial placentation may limit disease transmission between fetus and mother.

Author Manuscript Author Manuscript

As the nature of the maternal fetal interface evolves, so does the nature of the biochemical communication possible between conceptus and mother. This is particularly true in early pregnancy where there is an urgency to interrupt estrous cyclicity and maintain continued production of progesterone, if the pregnancy is to continue (Bazer, et al. 2009). As far as is known, it is invariably the trophoblast that is responsible for generating this signal for maternal recognition of pregnancy, although the process is understood in only a handful of species and even then rather poorly. What is clear is that the mechanisms are highly variable, evolving rapidly, and no one approach is used widely across taxa. A couple of examples illustrate the kind of strategies used in apes, where implantation occurs immediately after the blastocyst attaches, and in artiodactyls where implantation is superficial and delayed. In the human and most probably higher primates, but not in strepsirrhines (lemurs and related species) that have a non-invasive trophoblast, chorionic gonadotrophin (CG) provides luteotrophic support to the progesterone-producing cells of the corpus luteum soon after implantation has occurred. In women, hCG becomes measurable in blood and even urine by the second week of pregnancy, most likely as a result of being released directly into maternal capillaries even before the villous placental bed has formed (Roberts, et al. 1996). In ruminant species, the pregnant mother responds to the presence of the conceptus even before the trophoblast has attached to the uterine wall. The factor responsible is not a gonadotrophin, but instead a Type 1 interferon (IFN-tau; IFNT), which acts on the uterine epithelium, where it modulates the release of the luteolytic factor, prostaglandin F2α that would normally cause the corpus luteum to regress (Bazer, et al. 1997, Roberts, et al. 1999). In the pig, where maternal recognition of pregnancy is also achieved during the period when the conceptuses are elongating within the uterine lumen, the antiluteolytic factor appears not to be a protein, but is instead the steroid hormone estrogen (Spencer, et al. 2007). Selective mechanisms that may contribute to placental diversity—It is generally recognized that there are additional driving forces promoting adaptive changes in the placenta and hence contributing to its rapid evolution. One is undoubtedly the struggle for control of maternal physiology in terms of nutrient allocation (Fowden and Constancia 2012). As the demands of the fetus increase, they will likely conflict with the ability of the

Reproduction. Author manuscript; available in PMC 2017 November 01.

Roberts et al.

Page 8

Author Manuscript

mother to provide such resources. An extensive literature has developed around this topic, with particular emphasis on the role played by imprinted genes in controlling nutrient supply and growth of the fetus (Reik, et al. 2003). Many placental gene products, including hormones and transporters, for example, likely play a role in raising the concentrations of nutrients in maternal blood and facilitating their uptake by the placenta. There is, thus, a drive to keep the paternal genes representing the fetus active, while silencing the corresponding maternal alleles. Maternal efforts to counteract the acquisitiveness of the fetus may lead to the evolution of oppositely imprinted genes. Again, most speculation on the role of imprinted genes has come from the mouse, with roles inferred for the human, both species with an invasive placenta (Reik, et al. 2003). Very little is known about imprinting in those species where there is little or no direct access of the trophoblast to maternal blood, although data are beginning to emerge (Chen, et al. 2016).

Author Manuscript

A second force driving placental evolution is undoubtedly avoidance of immune rejection (Hemberger 2013), a process still little understood, although speculation abounds. It is assumed that to minimize attention from various arms of the maternal immune system, the trophoblast must continue to evolve counter-measures for its own protection and even advantage. These measures will undoubtedly be subtle and complex and change as the degree of intimacy with maternal system itself evolves. Some clues may emerge by examining candidates encoded by rapidly evolving genes and gene families within individual mammalian clades, some of which are discussed in the next section.

Evolution of placenta specific genes

Author Manuscript Author Manuscript

Despite the observation that the initial lineage regulators for trophoblast cell fate decisions may be at least partially conserved across mammals, the resulting placental structures and their associated trophoblast cell populations, appear not to be governed by particular master genes. There are a few coding genes, e.g. GCM1, a transcription factor that plays a role in controlling the formation of syncytial trophoblast (STB)(Cross, et al. 2003) that appear to have an entirely placental function in the sense that they have not been implicated in developmental events outside the trophoblast lineage. However, it seems likely that the much of the structural diversity and functional refinements associated with mammalian placentas depend on widely expressed transcription factors operating in a combinatorial, but trophoblast-specific manner and poorly understood processes linked to peculiarities in the epigenetic landscape of trophoblast. There have also beenlarge numbers of gene family expansions (Rawn and Cross 2008), These gene families may be performing roles in processes such as maternal recognition of pregnancy, nutrient acquisition, and immune protection, thereby contributing to the unique functional needs of different placental forms. It is worth noting that “placenta-specific” in this context means such genes are disproportionately, but not necessarily uniquely, expressed in the placenta. For example, it is not unusual for some placenta-specific gene products to be elevated in cancer and germ cells.

Reproduction. Author manuscript; available in PMC 2017 November 01.

Roberts et al.

Page 9

Placenta-specific gene families

Author Manuscript Author Manuscript

Gene duplication provides a useful opportunity for natural selection to operate (Hughes 1994) and allows the duplicated genes to acquire new functions, and, through the acquisition or loss of regulatory regions, changes in the temporal, spatial or magnitude of transcription compared to the original gene. Gene family expansions associated with distinct placental types have arisen repeatedly in eutherians (Rawn and Cross 2008). Here we provide four examples from the many known placenta-specific gene families. In general, assigning function of individual members within a family will be difficult. Indeed, knock-outs in mice performed on single members of gene families have generally not been informative, possibly because phenotype is not obviously impaired within the friendly confines of a vivarium, although it may become so when the pregnant animal is under stress (Ain, et al. 2004). Also products of multi-gene families likely overlap in function, so that a loss of one gene may be compensated by the presence of one or more of its orthologs. Finally, family members may even interact in an epistatic manner, providing a combinatorial benefit not achieved by a single member operating alone.

Author Manuscript

The growth hormone and prolactin-related proteins are found in primates, rodents, and ruminants and, most likely, other taxa. Interestingly, gene family expansions have occurred independently in these different lineages, providing yet another example of convergent evolution. Growth hormone (GH) and prolactin (PRL) themselves arose from a common ancestral gene. In most species the GH gene is single copy, but in humans, there are five members, one restricted in expression to the pituitary, the other four to the placenta (Haig 2008). Their likely role is in resource allocation between the fetus and the mother. By contrast, there is only a single copy of the PRL gene in the human, whereas there are 23 members in the mouse, 24 in the rat, and 12 in cattle (Haig 1993). All except prolactin itself are expressed in the placenta. In the mouse, they have diverse patterns of spatial and temporal expression (Simmons, et al. 2008) and appear to target the maternal endometrium (Soares, et al. 2007).

Author Manuscript

The pregnancy-specific glycoproteins (PSGs) are secreted molecules that are abundantly produced by primate and rodent placentas. They are related to the carcinoembryonic antigen cell adhesion molecules (CEACAM) found on the surface of certain tumors and selected normal somatic cells. There are 10 intact PSG genes in the human genome and 17 in mouse (McLellan, et al. 2005) although it is almost impossible to assign orthologous members based on sequence identity. Moreover, the organization of the domains in human and mouse PSG proteins are quite different (McLellan, et al. 2005), raising the possibility that their roles in primates and humans may not be identical. PSGs are able to enter the maternal circulation and, in the human, can accumulate to extraordinarily high concentrations in blood (200–400 μg/ml at term), but their roles remain enigmatic. The interferon tau (IFNT) proteins are products of the filamentous conceptus of ruminants and have been mentioned earlier in the context of maternal recognition of pregnancy. They have been described in no other taxa (Leaman and Roberts 1992). Within a species they are well conserved, but considerable divergence is evident in inter-species comparisons. They are transcribed from multiple, structurally conserved genes that have probably arisen by duplications and gene conversion events from another Type 1 IFN gene, interferon omega Reproduction. Author manuscript; available in PMC 2017 November 01.

Roberts et al.

Page 10

Author Manuscript

(IFNW). Although builds of the bovine genome have suggested the IFNT to be a relatively small gene family (Walker and Roberts 2009), large numbers of gene variants have been described in each species and even within single conceptuses, suggesting that the families continue to diverge rapidly. The IFNT proteins are secreted from mononucleated trophoblasts and act on the uterus by binding to type I interferon receptors, with the period of release occurring before the trophoblast has formed definite attachment to the uterine epithelium (Bazer, et al. 1997, Roberts, et al. 1999). Another notable feature of the IFNT is that these proteins have retained typical features of other type I interferons, such as potent antiviral activity, despite the fact that their primary role has shifted to a reproductive function. Conceivably, the progenitor gene product in the common ancestor of present day ruminants served to protect the conceptus from viral infections.

Author Manuscript

The pregnancy associated glycoproteins (PAGs) are products of a gene family present in placental trophoblasts of swine and ruminants and most other even-toed ungulates (Wallace, et al. 2015). They are related to aspartic proteinases, although some notably lack the capacity to cleave protein substrates due to modifications of normally conserved amino acids around the active sites. The PAGs are particularly complex in ruminants; in cattle, for example, there are approximately two dozen members along with some closely related paralogs. Most of the PAGs in ruminants are expressed by specialized polyploid trophoblast cells (binucleate giant cells) discussed earlier. After fusion, the contents of secretory granules within the binucleate cells are disgorged into the uterine stroma, and many of the released PAGs enter the maternal circulation and have formed the basis of useful pregnancy tests for various ruminants, including dairy cows. The role of PAGs is unknown, but their abundance suggests an important function during pregnancy.

Author Manuscript

Endogenous retroviruses and retroviral elements Ancestral retroviral infections have provided another source of novel protein-coding genes that have played a role in placental evolution. To ensure that the viral genes do not replicate, the control elements of endogenous retroviruses (ERVs) are usually highly methylated and effectively silenced. As a result, the fate of most ERVs is gradual genetic degradation through mutation and homologous recombination. Nevertheless, a minority of ERVs retain a whole or partially intact open reading frame potentially capable of producing a functional protein (Weiss 2016). This point is particularly salient in regard to the placenta, which in many species express a range of ERVs that are involved in trophoblast function (Denner 2016). Of those ERVs expressed in the placenta, the most studied have originated from the envelope (env) elements of the integrated viral DNA, and have been called syncytins (Lokossou, et al. 2014).

Author Manuscript

In humans, ERVW-1 (syncytin-1) facilitates fusion of mononucleated cytotrophoblasts to form multinucleated STBs (Mi, et al. 2000). This process involves the interaction of ERVW-1 with a “receptor”, most probably the neutral amino acid transporter, SLC5A1, on a neighboring cell. The expression of ERVW-1 is regulated by the placenta-specific transcription factor GCM1 discussed earlier, via an enhancer element present in the long terminal repeat (LTR) of the gene (Lin, et al. 2005). It is also notable that ERVW-1 is not involved in trophoblast fusion in Old World monkeys, despite being present in their genomes

Reproduction. Author manuscript; available in PMC 2017 November 01.

Roberts et al.

Page 11

Author Manuscript

(Cáceres, et al. 2006). These animals make use of a distinct syncytin, ERVV2, for STB formation. Conversely, ERVV2 is expressed in human placenta but does not appear to be involved in cytotrophoblast fusion (Blaise, et al. 2005). An important point illustrated here is that expressed retroviral envelope genes (env) often fulfill similar roles across species, but few represent orthologous genes. Instead, the majority has arisen as a result of independent integration events and coopted for similar functions by different species.

Author Manuscript

A second env gene from a different endogenous retrovirus (ERVFRD-1; syncytin-2) is expressed in villous trophoblast of the human placenta (Blaise, et al. 2003, Malassine, et al. 2007) and is also believed to be involved in trophoblast fusion events through interaction with another transporter protein, MFSD2. ERVFRD-1 has been proposed to be immunosuppressive, an activity not shared by ERVW-1 (Mangeney, et al. 2007). Several other ERV proteins, e.g.ERV3-1, and ERVK family members, including an antagonist of cell fusion (Sugimoto, et al. 2013) are expressed in the placenta, but roles for most are even more obscure than that of ERVFRD-1. In the mouse placenta the labyrinth layer, which is the functional equivalent of human floating villous trophoblast, forms the transport surface in contact with the maternal bloodstream. It is hemotrichorial, i.e. it has a cytotrophoblast layer overlaid by two STB layers (Dupressoir, et al. 2009). Two syncytin genes (Syna & Synb), products of separate integration events, are expressed in this tissue. The SYNB protein is primarily localized to the innermost STB layer, and, like ERVFRD in the human, may have immunosuppressive activity, as well as roles in facilitating cell fusion (Mangeney, et al. 2007). Neither of these gene products, however, is related to human syncytins, except for the fact they are derived from ERV genes and placentally expressed.

Author Manuscript

Finally, endogenous retroviruses, with no obvious homology to each other, have been described in the placentas of a number of other species, including marsupials, ruminants, carnivores, and lagomorphs (Denner 2016). In the ruminants they are mainly associated with trophoblast binucleate cells that fuse with uterine epithelial cells, but are assumed in all these species to play somewhat similar roles to syncytins in the human placenta. Conclusions

Author Manuscript

Placentas, whatever their anatomic features, share a number of common functions, all of which require cooperative interactions with the maternal system. Placentas deal with nutrient and oxygen import, dispose of fetal waste products and gases, help physically retain the conceptus in the reproductive tract, commandeer the local maternal blood supply in some manner, release factors, including metabolic hormones to adjust the needs of the fetus to the resources provided by the mother, and provide immune protection. Yet, despite these apparently conserved functions placentas are arguably the least conserved and most rapidly evolving mammalian organs. These changes are speculated to be driven, at least in part by maternal-fetal conflict (Haig 1996). Placental divergence has, in turn, been promoted by selection for multiple kinds of genetic change, including 1) duplications and gene conversion events to create large gene families that themselves are continuing to diverge extensively; 2) coopting of endogenous retrovirus-derived genes and gene control elements; 3) rapid evolution of placenta-specific enhancers and promoter elements; 4) imprinting. It is also Reproduction. Author manuscript; available in PMC 2017 November 01.

Roberts et al.

Page 12

Author Manuscript

possible that the hypomethylated state of the placental genome (Chuong 2013) has permitted relaxed silencing of gene control elements, including those of ERVs. Perhaps this feature of the placenta sets the stage for increased opportunity for epistasis, the interactions of genes that are not allelic. For example, a harmful mutation in one gene may occasionally become beneficial and confer increased fitness in the presence of a mutation in another gene (sign epistasis)(Weinreich, et al. 2005), possibly allowing both mutations to become fixed and the placenta to adapt to the constraints of a changing environment.

Acknowledgments We thank Dennis Reith for editorial help, Susan Roberts, Nico Zevallos, and Dr. Ye Yuan for assistance in preparing Figures Funding

Author Manuscript

This is supported by Missouri Mission Enhancement Funds (LCS), Food for the 21st Cnetury Support from the University of Missouri (RMR & JAG) and a NIH Eunice Kennedy Shriver National Institute of Child Health and Human Development Grant (R01 HD069979) (RMR).

References

Author Manuscript Author Manuscript

Ain R, Dai G, Dunmore JH, Godwin AR, Soares MJ. A prolactin family paralog regulates reproductive adaptations to a physiological stressor. Proc Natl Acad Sci U S A. 2004; 101:16543–16548. [PubMed: 15545614] Amoroso, E. Placentation. In: Parkes, A., editor. Marshall’s Physiology of Reproduction. Boston: Little Brown & Co.; 1952. p. 127-311. Bazer FW, Spencer TE, Johnson GA, Burghardt RC, Wu G. Comparative aspects of implantation. Reproduction. 2009; 138:195–209. [PubMed: 19502456] Bazer FW, Spencer TE, Ott TL. Interferon tau: a novel pregnancy recognition signal. Am J Reprod Immunol. 1997; 37:412–420. [PubMed: 9228295] Blackburn DG. Evolution of vertebrate viviparity and specializations for fetal nutrition: A quantitative and qualitative analysis. Journal of Morphology. 2015; 276:961–990. [PubMed: 24652663] Blackburn DG, Flemming AF. Invasive implantation and intimate placental associations in a placentotrophic african lizard, Trachylepis ivensi (scincidae). Journal of Morphology. 2012; 273:137–159. [PubMed: 21956253] Blaise S, de Parseval N, Bénit L, Heidmann T. Genomewide screening for fusogenic human endogenous retrovirus envelopes identifies syncytin 2, a gene conserved on primate evolution. Proceedings of the National Academy of Sciences. 2003; 100:13013–13018. Blaise S, de Parseval N, Heidmann T. Functional characterization of two newly identified Human Endogenous Retrovirus coding envelope genes. Retrovirology. 2005; 2:1–4. [PubMed: 15644139] Burton GJ, Jauniaux E. What is the placenta? Am J Obstet Gynecol. 2015; 213:S6–e1. S6–8. [PubMed: 26428504] Cáceres M, NCS Program. Thomas JW. The Gene of Retroviral Origin Syncytin 1 is Specific to Hominoids and is Inactive in Old World Monkeys. Journal of Heredity. 2006; 97:100–106. [PubMed: 16424151] Carcupino M, Baldacci A, Mazzini M, Franzoi P. Functional significance of the male brood pouch in the reproductive strategies of pipefishes and seahorses: a morphological and ultrastructural comparative study on three anatomically different pouches. Journal of Fish Biology. 2002; 61:1465–1480. Carter AM. Marsupial Frogs [Online]. 2016 posting date. Chen Z, Hagen DE, Wang J, Elsik CG, Ji T, Siqueira LG, Hansen PJ, Rivera RM. Global assessment of imprinted gene expression in the bovine conceptus by next generation sequencing. Epigenetics. 2016 0.

Reproduction. Author manuscript; available in PMC 2017 November 01.

Roberts et al.

Page 13

Author Manuscript Author Manuscript Author Manuscript Author Manuscript

Chuong EB. Retroviruses facilitate the rapid evolution of the mammalian placenta. BioEssays. 2013; 35:853–861. [PubMed: 23873343] Cornelis G, Vernochet C, Malicorne S, Souquere S, Tzika AC, Goodman SM, Catzeflis F, Robinson TJ, Milinkovitch MC, Pierron G, Heidmann O, Dupressoir A, Heidmann T. Retroviral envelope syncytin capture in an ancestrally diverged mammalian clade for placentation in the primitive Afrotherian tenrecs. Proc Natl Acad Sci U S A. 2014; 111:E4332–4341. [PubMed: 25267646] Cross JC, Baczyk D, Dobric N, Hemberger M, Hughes M, Simmons DG, Yamamoto H, Kingdom JC. Genes, development and evolution of the placenta. Placenta. 2003; 24:123–130. [PubMed: 12596737] Denner J. Expression and function of endogenous retroviruses in the placenta. APMIS. 2016; 124:31– 43. [PubMed: 26818260] Dupressoir A, Vernochet C, Bawa O, Harper F, Pierron G, Opolon P, Heidmann T. Syncytin-A knockout mice demonstrate the critical role in placentation of a fusogenic, endogenous retrovirusderived, envelope gene. Proceedings of the National Academy of Sciences. 2009; 106:12127– 12132. Enders AC, Carter AM. What can comparative studies of placental structure tell us?–A review. Placenta. 2004; 25(Suppl A):S3–9. [PubMed: 15033300] Familari M, Au PC, de Iongh RU, Cruz Y, Selwood L. Expression analysis of Cdx2 and Pou5f1 in a marsupial, the stripe-faced dunnart, during early development. Mol Reprod Dev. 2016; 83:108– 123. [PubMed: 26613191] Fowden AL, Constancia M. Maternal-fetal resource allocation. Placenta. 2012; 33(Suppl 2):e1–2. [PubMed: 23000397] Frankenberg S, Shaw G, Freyer C, Pask AJ, Renfree MB. Early cell lineage specification in a marsupial: a case for diverse mechanisms among mammals. Development. 2013; 140:965–975. [PubMed: 23344710] Freyer C, Zeller U, Renfree MB. The marsupial placenta: A phylogenetic analysis. Journal of Experimental Zoology Part a-Comparative Experimental Biology. 2003; 299A:59–77. Friess AE, Sinowatz F, Skolek-Winnisch R, Traautner W. The placenta of the pig. I. Finestructural changes of the placental barrier during pregnancy. Anat Embryol. 1980; 158:179–191. [PubMed: 7356176] Friess AE, Sinowatz F, Skolek-Winnisch R, Trautner W. The placenta of the pig. II. The ultrastructure of the areolae. Anat Embryol (Berl). 1981; 163:43–53. [PubMed: 7316222] Haig D. Genetic conflicts in human pregnancy. Q Rev Biol. 1993; 68:495–532. [PubMed: 8115596] Haig D. Altercation of generations: genetic conflicts of pregnancy. Am J Reprod Immunol. 1996; 35:226–232. [PubMed: 8962651] Haig D. Placental growth hormone-related proteins and prolactin-related proteins. Placenta. 2008; 29(Suppl A):S36–41. [PubMed: 17981323] Hamlett WC, Eulitt AM, Jarrell RL, Kelly MA. Uterogestation and Placentation in Elasmobranchs. Journal of Experimental Zoology. 1993; 266:347–367. Hemberger M. Immune balance at the foeto-maternal interface as the fulcrum of reproductive success. J Reprod Immunol. 2013; 97:36–42. [PubMed: 23432870] Hubrecht AA. The Trophoblast: A Rejoinder. Science. 1904; 20:367–370. [PubMed: 17734040] Hughes AL. The evolution of functionally novel proteins after gene duplication. Proc Biol Sci. 1994; 256:119–124. [PubMed: 8029240] Kalinka AT. How did viviparity originate and evolve? Of conflict, co-option, and cryptic choice. BioEssays. 2015; 37:721–731. [PubMed: 25904118] Korsgaard, B.; Weber, RE. Maternal-Fetal Trophic and Respiraotry Relationships. Berlin ; New York: Springer-Verlag; 1989. Kwan L, Fris M, Rodd FH, Rowe L, Tuhela L, Panhuis TM. An examination of the variation in maternal placentae across the genus Poeciliopsis (Poeciliidae). J Morphol. 2015; 276:707–720. [PubMed: 25765517]

Reproduction. Author manuscript; available in PMC 2017 November 01.

Roberts et al.

Page 14

Author Manuscript Author Manuscript Author Manuscript Author Manuscript

Leaman DW, Roberts RM. Genes for the trophoblast interferons in sheep, goat, and musk ox and distribution of related genes among mammals. J Interferon Res. 1992; 12:1–11. [PubMed: 1374107] Lin C, Lin M, Chen H. Biochemical characterization of the human placental transcription factor GCMa/1. Biochemistry and Cell Biology. 2005; 83:188–195. [PubMed: 15864327] Lokossou A, Toudic C, Barbeau B. Implication of Human Endogenous Retrovirus Envelope Proteins in Placental Functions. Viruses. 2014; 6:4609. [PubMed: 25421890] Malassine A, Blaise S, Handschuh K, Lalucque H, Dupressoir A, Evain-Brion D, Heidmann T. Expression of the fusogenic HERV-FRD Env glycoprotein (syncytin 2) in human placenta is restricted to villous cytotrophoblastic cells. Placenta. 2007; 28:185–191. [PubMed: 16714059] Mangeney M, Renard M, Schlecht-Louf G, Bouallaga I, Heidmann O, Letzelter C, Richaud A, Ducos B, Heidmann T. Placental syncytins: Genetic disjunction between the fusogenic and immunosuppressive activity of retroviral envelope proteins. Proceedings of the National Academy of Sciences. 2007; 104:20534–20539. McLellan AS, Zimmermann W, Moore T. Conservation of pregnancy-specific glycoprotein (PSG) N domains following independent expansions of the gene families in rodents and primates. BMC Evolutionary Biology. 2005; 5:1–19. [PubMed: 15629062] Mess A, Carter AM. Evolution of the placenta during the early radiation of placental mammals. Comp Biochem Physiol A Mol Integr Physiol. 2007; 148:769–779. [PubMed: 17347003] Mi S, Lee X, Li X-p, Veldman GM, Finnerty H, Racie L, LaVallie E, Tang X-Y, Edouard P, Howes S, Keith JC, McCoy JM. Syncytin is a captive retroviral envelope protein involved in human placental morphogenesis. 2000; 403:785–789. Morrison JT, Bantilan NS, Wang VN, Nellett KM, Cruz YP. Expression patterns of Oct4, Cdx2, Tead4, and Yap1 proteins during blastocyst formation in embryos of the marsupial, Monodelphis domestica Wagner. Evol Dev. 2013; 15:171–185. [PubMed: 23607301] Murphy BF, Parker SL, Murphy CR, Thompson MB. Placentation in the eastern water skink (Eulamprus quoyii): a placentome-like structure in a lecithotrophic lizard. Journal of Anatomy. 2011; 218:678–689. [PubMed: 21434912] Painter DL, Moore MC. Steroid hormone metabolism by the chorioallantoic placenta of the mountain spiny lizard Sceloporus jarrovi as a possible mechanism for buffering maternal-fetal hormone exchange. Physiological and Biochemical Zoology. 2005; 78:364–372. [PubMed: 15887083] Pollux BJ, Meredith RW, Springer MS, Garland T, Reznick DN. The evolution of the placenta drives a shift in sexual selection in livebearing fish. Nature. 2014; 513:233–236. [PubMed: 25043015] Posfai E, Tam OH, Rossant J. Mechanisms of pluripotency in vivo and in vitro. Curr Top Dev Biol. 2014; 107:1–37. [PubMed: 24439801] Pyron RA, Burbrink FT. Early origin of viviparity and multiple reversions to oviparity in squamate reptiles. Ecology Letters. 2014; 17:13–21. [PubMed: 23953272] Rawn SM, Cross JC. The Evolution, Regulation, and Function of Placenta-Specific Genes. Annual Review of Cell and Developmental Biology. 2008; 24:159–181. Reik W, Constancia M, Fowden A, Anderson N, Dean W, Ferguson-Smith A, Tycko B, Sibley C. Regulation of supply and demand for maternal nutrients in mammals by imprinted genes. J Physiol. 2003; 547:35–44. [PubMed: 12562908] Renegar RH, Bazer FW, Roberts RM. Placental Transport and Distribution of Uteroferrin in the Fetal Pig. Biol Reprod. 1982; 27:1247–1260. [PubMed: 7159665] Renfree, MB. Implantation and Placentation Vol 2, Reproduction in Mammals: Embryonic and Fetal Development. Austin, CR.; Short, RV., editors. Cambridge University Press; Cambridge: 1982. p. 26-69. Renfree MB. Review: Marsupials: placental mammals with a difference. Placenta. 2010; 31(Suppl):S21–26. [PubMed: 20079531] Roberts RM, Ealy AD, Alexenko AP, Han CS, Ezashi T. Trophoblast interferons. Placenta. 1999; 20:259–264. [PubMed: 10329345] Roberts RM, Xie S, Mathialagan N. Maternal recognition of pregnancy. Biol Reprod. 1996; 54:294– 302. [PubMed: 8788179]

Reproduction. Author manuscript; available in PMC 2017 November 01.

Roberts et al.

Page 15

Author Manuscript Author Manuscript Author Manuscript Author Manuscript

Savage, JM. The amphibians and reptiles of Costa Rica : a herpetofauna between two continents, between two seas. Chicago: University of Chicago Press; 2002. Selwood L, Johnson MH. Trophoblast and hypoblast in the monotreme, marsupial and eutherian mammal: evolution and origins. BioEssays. 2006; 28:128–145. [PubMed: 16435291] Simmons DG, Rawn S, Davies A, Hughes M, Cross JC. Spatial and temporal expression of the 23 murine Prolactin/Placental Lactogen-related genes is not associated with their position in the locus. BMC Genomics. 2008; 9:352. [PubMed: 18662396] Soares MJ, Konno T, Alam SMK. The prolactin family: effectors of pregnancy-dependent adaptations. Trends in Endocrinology & Metabolism. 2007; 18:114–121. [PubMed: 17324580] Spencer TE, Johnson GA, Bazer FW, Burghardt RC, Palmarini M. Pregnancy recognition and conceptus implantation in domestic ruminants: roles of progesterone, interferons and endogenous retroviruses. Reprod Fertil Dev. 2007; 19:65–78. [PubMed: 17389136] Stewart JR. Placental specializations in lecithotrophic viviparous squamate reptiles. Journal of Experimental Zoology Part B-Molecular and Developmental Evolution. 2015; 324:549–561. Sugimoto J, Sugimoto M, Bernstein H, Jinno Y, Schust D. A novel human endogenous retroviral protein inhibits cell-cell fusion. Sci Rep. 2013; 3:1462. [PubMed: 23492904] Thompson MB, Speake BK. Energy and nutrient utilisation by embryonic reptiles. Comparative Biochemistry and Physiology a-Molecular & Integrative Physiology. 2002; 133:529–538. Van Dyke JU, Brandley MC, Thompson MB. The evolution of viviparity: molecular and genomic data from squamate reptiles advance understanding of live birth in amniotes. Reproduction. 2014; 147:R15–26. [PubMed: 24129151] Wake MH. Evolution of Oviductal Gestation in Amphibians. Journal of Experimental Zoology. 1993; 266:394–413. Walker AM, Roberts RM. Characterization of the bovine type I IFN locus: rearrangements, expansions, and novel subfamilies. BMC Genomics. 2009; 10:1–15. [PubMed: 19121221] Wallace RM, Pohler KG, Smith MF, Green JA. Placental PAGs: gene origins, expression patterns, and use as markers of pregnancy. Reproduction. 2015; 149:R115–R126. [PubMed: 25661256] Weekes HC. A review ofplacentation among reptiles with particular regard tofunction and evolution of the placenta. Proceedings of the Zoological Society of London. 1935; 2:625–640. Weinreich DM, Watson RA, Chao L. Perspective: Sign epistasis and genetic constraint on evolutionary trajectories. Evolution. 2005; 59:1165–1174. [PubMed: 16050094] Weiss RA. Human endogenous retroviruses: friend or foe? APMIS. 2016; 124:4–10. [PubMed: 26818257] Wildman DE, Chen C, Erez O, Grossman LI, Goodman M, Romero R. Evolution of the mammalian placenta revealed by phylogenetic analysis. Proc Natl Acad Sci U S A. 2006; 103:3203–3208. [PubMed: 16492730] Wilson AB, Vincent A, Ahnesjo I, Meyer A. Male pregnancy in seahorses and pipefishes (Family Syngnathidae): Rapid diversification of paternal brood pouch morphology inferred from a molecular phylogeny. Journal of Heredity. 2001; 92:159–0166. [PubMed: 11396574] Wooding FB. Current topic: the synepitheliochorial placenta of ruminants: binucleate cell fusions and hormone production. Placenta. 1992; 13:101–113. [PubMed: 1631024] Wourms JP, Lombardi J. Reflections on the Evolution of Piscine Viviparity. American Zoologist. 1992; 32:276–293. Zeller U, Freyer C. Early ontogeny and placentation of the grey short-tailed opossum, Monodelphis domestica (Didelphidae : Marsupialia): contribution to the reconstruction of the marsupial morphotype. Journal of Zoological Systematics and Evolutionary Research. 2001; 39:137–158.

Reproduction. Author manuscript; available in PMC 2017 November 01.

Roberts et al.

Page 16

Author Manuscript Author Manuscript

Figure 1.

Author Manuscript

Viviparity in reptiles: A, Drawing from a section of the mature chorio-allantoic placenta of the Australian skink Egernia cunninghami carrying developing young, showing the maternal capillaries closely adjoining the allantoic capillaries on the fetal side. B, Drawing from a section of the immature chorio-allantoic placenta of the Australian skink Egernia entrecasteaux. The section illustrates the folded placental face likely involved in releasing histotrophic material that can be taken up by the chorionic ectoderm. Note also the close apposition of maternal blood capillaries with the epithelium of the reproductive tract. The diagrams are based on Fig. 15.5 and 15.7 from E.C. Amoroso’s review on placentation (Amoroso 1952), which were redrawn from Weekes (Weekes 1935).

Author Manuscript Reproduction. Author manuscript; available in PMC 2017 November 01.

Roberts et al.

Page 17

Author Manuscript Author Manuscript Figure 2.

Cartoon illustrating the diversity of placental morphologies encountered in placental mammals

Author Manuscript Author Manuscript Reproduction. Author manuscript; available in PMC 2017 November 01.

Roberts et al.

Page 18

Author Manuscript Author Manuscript

Figure 3.

Cartoon indicating how placentas can be classified according to the numbers and kinds of cell layers that separate the bloodstreams of the mother and conceptus

Author Manuscript Author Manuscript Reproduction. Author manuscript; available in PMC 2017 November 01.

Roberts et al.

Page 19

Author Manuscript Author Manuscript Author Manuscript

Figure 4.

The evolution of the placental interface in terms of degree of invasiveness of the placental tissue into maternal tissue, with epitheliochorial being the least invasive and hemochorial being most invasive. From Fig. 2 of Wildman et al. (Wildman, et al. 2006), with permission; Copyright (2006) National Academy of Sciences, U.S.A.

Author Manuscript Reproduction. Author manuscript; available in PMC 2017 November 01.

Roberts et al.

Page 20

Author Manuscript Author Manuscript Author Manuscript

Figure 5.

Author Manuscript

The porcine placenta: A, Fetal capillaries (FC) at d 58 of pregnancy protrude deeply into the trophoblast, with the diffusion distance between fetal bloodstream and maternal uterine epithelium (UE) reduced to a few μm; B, At d 110 of pregnancy maternal capillaries (MC) project between uterine epithelial cells bringing the maternal and fetal capillaries within 3–5 μm; C, At d 30 of pregnancy, the microvilli on the trophoblast surface (TR) interdigitate with ones on the uterine epithelium (UE) to provide an intimate contact layer. Maternal capillaries (MC) are placed just below the basal lamina of the UE; D, A general view of a dome-shaped areola (AE) situated above the mouth of a uterine gland (UG) at d 30 of pregnancy. Figures 5A-C are from Friess et al. (1980) (Friess, et al. 1980); Figure 5D is from Friess et al. (1981)(Friess, et al. 1981) with permission.

Reproduction. Author manuscript; available in PMC 2017 November 01.

The evolution of the placenta.

The very apt definition of a placenta is coined by Mossman, namely apposition or fusion of the fetal membranes to the uterine mucosa for physiological...
1MB Sizes 0 Downloads 9 Views