Next Article in Journal
The Genetic Landscape of Parkinsonism-Related Dystonias and Atypical Parkinsonism-Related Syndromes
Next Article in Special Issue
CSH RNA Interference Reduces Global Nutrient Uptake and Umbilical Blood Flow Resulting in Intrauterine Growth Restriction
Previous Article in Journal
Super-Toughened Fumed-Silica-Reinforced Thiol-Epoxy Composites Containing Epoxide-Terminated Polydimethylsiloxanes
Previous Article in Special Issue
Immunoendocrine Dysregulation during Gestational Diabetes Mellitus: The Central Role of the Placenta
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Unique Aspects of Human Placentation

by
Anthony M. Carter
Cardiovascular and Renal Research, Institute of Molecular Medicine, University of Southern Denmark, DK-5230 Odense, Denmark
Int. J. Mol. Sci. 2021, 22(15), 8099; https://doi.org/10.3390/ijms22158099
Submission received: 30 June 2021 / Revised: 21 July 2021 / Accepted: 26 July 2021 / Published: 28 July 2021
(This article belongs to the Special Issue Placental Related Disorders of Pregnancy)

Abstract

:
Human placentation differs from that of other mammals. A suite of characteristics is shared with haplorrhine primates, including early development of the embryonic membranes and placental hormones such as chorionic gonadotrophin and placental lactogen. A comparable architecture of the intervillous space is found only in Old World monkeys and apes. The routes of trophoblast invasion and the precise role of extravillous trophoblast in uterine artery transformation is similar in chimpanzee and gorilla. Extended parental care is shared with the great apes, and though human babies are rather helpless at birth, they are well developed (precocial) in other respects. Primates and rodents last shared a common ancestor in the Cretaceous period, and their placentation has evolved independently for some 80 million years. This is reflected in many aspects of their placentation. Some apparent resemblances such as interstitial implantation and placental lactogens are the result of convergent evolution. For rodent models such as the mouse, the differences are compounded by short gestations leading to the delivery of poorly developed (altricial) young.

1. Introduction

Adverse pregnancy outcomes can often be linked to defects in placentation [1]. Ethical considerations preclude detailed exploration of the underlying mechanisms. Unfortunately, there are also limitations to what can be learned from animal models. The mouse (Mus musculus) and other murine rodents have exceedingly short gestations. Whilst they may be informative about early events, such as the differentiation of cell lineages, they are unsatisfactory for modelling the events of third-trimester human pregnancy [2]. In addition, there are important differences between rodent and human in placentation and the disposition of fetal membranes such as the yolk sac. The objective of this review is to discuss these unique features of human placentation, to define their appearance during the evolution of primates and to contrast them with rodents.
Primary functions of the placenta are gas exchange and the transfer of substrates from mother to fetus. The underlying mechanisms are similar across mammals. Thus the sheep is an excellent model for studying the oxygen supply to the fetus despite structural differences between human and ovine placentation [3]. Similarly, glucose transfer by facilitated diffusion uses the same set of transporters across species [4]. These topics will not be further explored. The interactions between the trophoblast and the maternal immune system are manifold, and a full reckoning cannot be made here. The section on placental immunology therefore focuses mainly on the uterine natural killer (uNK) cells and their ligands. Information on other immune cells, including macrophages, T-cells and innate lymphoid cells, should be sought elsewhere [5,6].

Mammalian Evolution and Phylogeny

The uniqueness of human placentation can best be assessed in the evolutionary framework provided by phylogenetics. I have striven to keep terminology to a minimum, yet some context is needed, especially for primates. In a broader perspective, eutherian mammals can be sorted into four major clades (Figure 1A). Here we shall deal mainly with one of those clades (Euarchontoglires) and its two subdivisions (Figure 1B). Euarchonta comprises primates, tree shrews and colugos. Glires comprises rodents and lagomorphs. The split between Euarchonta and Glires is estimated to have occurred in the Cretaceous period some 80 million years ago [7]. Therefore, it is not surprising that placentation in the mouse and other rodent models differs in significant respects from human placentation [2].
The primate order has two major subdivisions: Strepsirrhini and Haplorrhini (Figure 2). The former comprises lemurs and lorises with placentation that differs radically from that of humans [10]. In contrast, Haplorrhini, to which our species belongs, was defined by commonalities in fetal membrane development [11]. It includes tarsiers (Tarsiiformes), New World monkeys (Platyrrhini), Old World monkeys and apes (together Catarrhini). Gibbons and great apes (orang-utans, gorillas, bonobo, chimpanzees, and man) comprise the superfamily Hominoidea.

2. Early Development

2.1. Interstitial Implantation

In most primates, implantation of the blastocyst is superficial. In macaques and baboons, for example, the trophoblast invades the endometrium to establish a placenta, but the developing embryo remains in the uterine cavity. In contrast, the human blastocyst is pulled into the endometrium, which closes above it so that it is completely embedded by the 12th day [13]. The placental bed is underlain by the basal decidua, and the developing embryo is covered by the capsular decidua. Interstitial implantation is a feature shared with the great apes and the gibbons [14,15]. It does occur in rodents, but the process is not identical and has been independently evolved.
Initial penetration of the endometrium is achieved by syncytiotrophoblast [16]. This is formed by fusion of cellular trophoblast to form a multicellular syncytium. The process depends in large part upon syncytins, which are proteins encoded by endogenous retroviral envelope genes that have been incorporated in the genome and exapted to promote cell fusion in the placenta [17]. Humans have two syncytin genes acquired at different timepoints. Whereas Syncytin-2 occurs in all haplorrhine primates, Syncytin-1 is found only in apes [17]. Syncytin genes occur in a wide range of mammals, and each represents a separate gene capture [17]. However, it has been argued that the capture of retroviral envelope genes was a prerequisite for the evolution of invasive placentation in mammals [18].

2.2. Initial Decidual Reaction

The maternal response to implantation is the decidual reaction, which involves the transformation of fibroblast-like endometrial stromal cells into polygonal decidual stromal cells [19]. The decidual reaction once was thought to be absent or atypical in elephants and carnivores [20], yet recent work shows it to be a characteristic feature of eutherian mammals [19]. However, the decidual reaction is transient in some species, such as the nine-banded armadillo (Dasypus novemcinctus), and has been lost in many with non-invasive placentation such as cattle (Bos taurus) [21]. It is now thought that the decidual reaction evolved from an inflammatory response that is present in marsupials, where it imposes a limit on the length of gestation [22,23]. Several of the genes involved in this response have been downregulated in eutherians, while genes beneficial to implantation have been upregulated [24].
In humans and many primates, as well as rodents, decidual stromal cells persist throughout gestation and have acquired an additional role in pregnancy maintenance [19]. These novel functions arose in the lineage of the large clade Euarchontoglires [19].

2.3. Early Differentiation of Mesoderm and Secondary Yolk Sac

One of the first fetal membranes to form in mammals is a bilaminar yolk sac comprising an outer layer of trophoblast and an inner lining of the extraembryonic endoderm. It may later acquire blood vessels and function as a choriovitelline placenta. In humans, however, the primary yolk sac is short-lived due to precocious differentiation of the extraembryonic mesoderm, which intrudes between the endoderm and trophoblast (Figure 3). This leads to formation of the secondary yolk sac, which consists of mesoderm and endoderm and becomes a free floating structure within the exocoelomic cavity [25]. Despite lack of contact with maternal tissues, the secondary yolk sac plays an important role in nutrient supply to the first trimester embryo [26]. Precocious development of the extraembryonic mesoderm is a defining feature of haplorrhine primates [11]. Recent work comparing gene expression in the common marmoset (Callithrix jacchus), rhesus macaque (Macaca mulatta) and human suggests extraembryonic mesoderm is derived in part from the extraembryonic endoderm [27,28].
Rodents pursue an entirely different course resulting in an inverted yolk sac with an outward-facing layer of endoderm that persists throughout pregnancy [29]. There are similarities but also marked differences in gene expression and regulatory pathways between the mouse and primates [28].

2.4. Allantoic Stalk

The chorioallantoic placenta is formed by the fusion of the allantois with the chorion (trophoblast and extraembryonic mesoderm). In most mammals, the allantois also encloses a fluid-filled space. Indeed, a medium to large allantoic sac is the ancestral state for eutherians [30], and it forms a prominent structure in some species. As an example, cattle have 6–9 litres of allantoic fluid against 2.5 litres of amniotic fluid [31]. In contrast, the human allantois develops as a small diverticulum, and the connection between embryo and placenta, which carries the blood vessels, is the allantoic stalk. This is yet another shared feature that defines haplorrhine primates [11,28]. An allantoic sac is absent in rodents, but this may be due to convergent evolution as it is found in their sister group, the lagomorphs (e.g., rabbit Oryctolagus cuniculus) [29].

3. Placentation

3.1. Haemochorial Placentation

Invasive placentation is the basal condition in eutherians [30,32]. Some strepsirrhine primates (lemurs and lorises) have epitheliochorial placentation, but this is widely regarded as a derived trait [10]. In haemochorial placentas, the trophoblast is in direct contact with maternal blood. Early in human gestation the interhaemal barrier includes two layers of trophoblast [33], but the cytotrophoblast layer (Langhan’s layer) later becomes discontinuous. The interhaemal barrier then comprises syncytiotrophoblast, a thin layer of connective tissue, and the fetal capillary endothelium (Figure 4). Therefore, human placenta is classified as haemomonochorial [34]. Mouse and rat (Rattus norvegicus) have three layers of trophoblast, rabbits two, and guinea pigs (Cavia porcellus) one, but the significance of these differences should not be overstated [2].

3.2. Villous rather Than Labyrinthine Placentation

Of greater consequence is the internal structure of the placenta. Most haemochorial placentas are labyrinthine and organised so that maternal blood channels are arranged in parallel with fetal capillaries. Maternal and fetal blood flow in opposite directions allowing for efficient countercurrent exchange [36,37]. This is also the case in the mouse [38]. In human placenta, on the other hand, the terminal villi are suspended in the intervillous space, which is supplied with blood by the spiral arteries of the basal plate. This is a pattern shared with Old World monkeys and apes [39]. An intermediate form of placentation, where the villi remain connected by bridges of trophoblast (trabeculae), is found in tarsiers and New World monkeys. The evolution of a villous placenta from a labyrinthine one negates the benefit conferred by countercurrent exchange. However, opening up the maternal component allows for a larger volume flow of blood and, thus, greater oxygen delivery, and this likely outweighs the loss of countercurrent exchange [4].

3.3. Uterine Spiral Artery Transformation

A key feature of human placentation is the transformation of the uterine spiral arteries to wide vessels with low resistance to flow. An initial phase involves vacuolation and focal loss of endothelial cells and loosening of the smooth muscle layer. It appears to be dependent on cytokines secreted by uNK cells and macrophages [40,41,42]. The second phase is associated with invasion of the endometrium and vessel walls by extravillous trophoblast. This leads to complete loss of endothelium and disruption of the smooth muscle with the greatly widened vessels eventually being lined by trophoblast embedded in a fibrinoid layer [41].
With advancing pregnancy there is also dilatation of the radial arteries, arcuate arteries and uterine arteries, none of which are invaded by trophoblast. This is most likely due to stimulation by oestrogens and nitric oxide-mediated flow-dilation signals [43].

3.4. Trophoblast Invasion by Interstitial and Intravascular Routes

In human pregnancy, the trophoblast invades by two routes [16]. Firstly, it migrates from the basal plate into the lumina of the uterine spiral arteries against the direction of flow (the intravascular route). Secondly, the trophoblast differentiating from the anchoring villi migrates through the decidua towards the blood vessels (the interstitial route). In a healthy pregnancy, trophoblast invasion extends through the relatively shallow endometrium to the inner third of the myometrium. Shallower invasion leads to inadequate transformation of the spiral arteries, thereby limiting the blood supply to the intervillous space, and is causally associated with fetal growth restriction and preeclampsia [1,44]. Trophoblasts that invade by the interstitial route undergo endoreduplication [45] and go no deeper than the inner myometrium, where they are found as multinucleate giant cells [46].
Our studies in chimpanzee and gorilla suggest that the depth of trophoblast invasion and spiral artery transformation resembles the human condition [47,48]. In Old World monkeys, however, there is rapid invasion of spiral arteries by the intravascular route but none by the interstitial route; indeed, there is a sharp border between the cytotrophoblastic shell and decidua [49,50]. This also appears to be the case in gibbons, suggesting that invasion by the interstitial route evolved in the lineage of the great apes [51]. In Old World monkeys, intravascular trophoblast is confined largely to the endometrial segments of the spiral arteries [49]. The situation in New World monkeys and tarsiers is not sufficiently known [16].
Rodent models do not readily conform to any of these features. The depth of trophoblast invasion in rodents varies but can extend to the mesometrial arteries, as in the guinea pig [52]. In the mouse, trophoblast glycogen cells migrate to the decidua but do not invade its vessels [53]. Instead, trophoblasts of the giant cell lineage migrate to and line the spiral arteries [54]. However, this does not occur until vascular remodelling is complete [5,55]. Indeed, rodents are unsatisfactory models of trophoblast invasion. Thus, deeper penetration of the arteries was found in a rat model of preeclampsia [56], which is the opposite of the shallower invasion typical of preeclampsia in human pregnancy [1].

4. Immunology of Decidua and Trophoblast

The placenta is a semi-allograft, yet it is not rejected by the maternal immune system. This immunological paradox was framed by Sir Peter Medawar [57] and remains a pivotal question in reproductive immunology [6]. As many as 70% of leukocytes in the uterus are uNK cells [58]. Their properties differ from those of peripheral natural killer cells and include lower cytotoxic activity [42]. Their putative role in the early stages of uterine artery transformation, alluded to above, may be explained by the secretion of cytokines, growth factors and proteases [40,42]. Here, we are concerned with their interplay with invasive extravillous trophoblast.

Interplay of uNK Cell Receptors and HLA Antigens

Human trophoblast does not express the major histocompatibility antigens (MHC) Class I, which include human leukocyte antigens (HLA) A and B. Instead, the surface of trophoblast presents HLA-C, which exhibits a high degree of allelic polymorphism, as well as HLA-E and HLA-G. There is no equivalent to HLA-C in monkeys or gibbons [59,60], but an invariant form appears in orangutans [61]. A later gene duplication yielded the two epitopes C1 and C2, which are found in chimpanzees, gorillas and humans [62].
HLA-C1 and -C2 are the principal ligands for the killer immunoglobulin-like receptors (KIRs) on uNK cells. KIRs likewise exhibit a high degree of allelic polymorphism, and importantly, there are inhibitory and excitatory variants [63]. Thus, numerous combinations are possible of HLA-C presented by the trophoblast (with one allele being paternal in origin) and KIRs expressed by maternal uNK cells. This can affect pregnancy outcome. When the trophoblast expresses HLA-C2 and the uNK cells express only inhibitory receptors, the combination is associated with a higher incidence of recurrent abortion, fetal growth restriction and preeclampsia [64,65]. KIR genes have evolved along separate pathways in great apes and human [66]. Indeed, it has been proposed that the emergence of the HLA-C2 epitope in apes is causally linked to the advent of preeclampsia. This is difficult to prove as reports of eclampsia in great apes are largely anecdotal (see [51]). In any case, the evolution of KIRs has pursued different paths in nonhuman primates [67].
HLA-G is expressed exclusively on trophoblast and has been implicated in maternal immune tolerance [68,69]. MHC-G is expressed in the great apes [70] but is a pseudogene in baboons (Papio spp.), rhesus macaque (Macaca mulatta), cynomolgus macaque (M. fascicularis) and vervet monkey (Chlorocebus aethiops), where its function is assumed by a new gene MHC-AG [60,71]. Receptors for HLA-G and MHA-AG are expressed by uNK cells and include KIR2DL4 [60].
The uterus of rodents is also rich in uNK cells, and they appear to be essential for the transformation of vessels analogous to human spiral arteries [72,73]. However, the principal receptors on rodent uNK cells belong to the lectin-like family (Ly49) [74], so rodents are not useful for exploring interactions between KIRs and HLA antigens. Rodents do not have MHC-G, although HLA-G expression has been achieved in transgenic mice [75].

5. Endocrinology of the Placenta

Placental hormones are secreted to the maternal circulation and adapt maternal physiology to meet the requirements of pregnancy and subsequent lactation [76]. Many of the peptides made by human trophoblast are unique to the primate lineage. Placental lactogens occur in human and rodents but have arisen through convergent evolution and serve different functions. Pregnancy maintenance depends on progesterone secretion, but an unusual feature of human pregnancy is that parturition occurs without a fall in plasma progesterone.

5.1. Chorionic Gonadotrophins

Human chorionic gonadotrophin (hCG) is responsible for early pregnancy maintenance. It evolved through duplication of the gene encoding the β-subunit of luteinising hormone. This occurred in the lineage of haplorrhine primates followed by further duplications so that many primates have multiple genes and pseudogenes [77]. A chorionic gonadotrophin was convergently evolved in the lineage of equids [78].

5.2. Placental Lactogens and Growth Hormones

Human placenta also secretes a placental lactogen, hPL, previously known as chorionic somatomammotropin. The genes that code for hPL are derived from the growth hormone gene [79]. A third gene in the cluster codes for placental growth hormone, which supplants pituitary growth hormone in the latter part of pregnancy [80]. All haplorrhine primates have placentally expressed genes related to growth hormone but there is great variation especially between New World and Old World monkeys [81,82], which may have attained placental expression separately [83].
The placental lactogens of muroid rodents, PL1 and PL2, are responsible for the maintenance of the corpus luteum [84]. In contrast to primates, they were derived by duplication from the prolactin gene together with a range of other cytokines [84]. Thus, placental lactogens of primates and rodents differ both in origin and function.

5.3. Progesterone and Its Receptors

Pregnancy maintenance in mammals requires the presence of progesterone secreted from the corpus luteum or placenta [85]. In many species, parturition is initiated through a fall in circulating progesterone, so-called progesterone withdrawal. In humans, where the placenta synthesises progesterone from maternal cholesterol, secretion is maintained right up to the start of labour [86]. This contra-intuitive finding led to the concept of a “functional” progesterone withdrawal for which the favoured explanation focusses on progesterone receptors (PR) in the myometrium. Of the two major isoforms, PR-B is the stronger trans-activator of progesterone-responsive genes, and PR-A acts as a trans-suppressor of PR-B’s effect. They are equally expressed in the myometrium throughout gestation. However, parturition is associated with a change in the PR-A to PR-B ratio due to increased expression of PR-A [87]. This switch contributes to myometrial activation via the activator protein-1 (AP-1) pathway [88]. Interestingly, there is evidence for adaptive evolution of the progesterone receptor gene (PGR) in the human lineage [89,90]. Of note, PRs are also expressed in human decidua, and a recent hypothesis points to a decline in decidual PR expression as a possible factor in the initiation of parturition [91].
Plasma progesterone levels increase before parturition in the rhesus macaque [92], and there is evidence of rapid evolution of PGR in catarrhine primates [90].
In mouse and rat, the corpus luteum is the sole source of progesterone, and plasma concentrations fall precipitously before parturition. These models are of limited value in understanding human parturition [93]. The situation is different in hystricomorph rodents: the placenta is a major source of progesterone in the guinea pig, and there is no change in circulating progesterone prior to parturition [86]. Therefore, it is considered a more appropriate model for parturition research [93].

6. Pregnancy Duration and Newborn State

An undeniably unique feature of human reproduction is that newborn babies are helpless and entirely dependent on parental care [94]. They differ to some degree from other haplorrhine primates, although a long childhood is a general feature of great apes (Table 1).
Mammals tend to adopt one of two contrasting strategies [108]. In the first, a short gestation with a large litter leads to the birth of poorly developed or altricial offspring. In the second, a long gestation with a small litter (usually singleton) ends with the birth of well-developed offspring with open eyes and ears, a coat of hair, and some degree of independence. Human babies have most of the attributes of precocial offspring but are helpless at birth and require parental care for several years.
The human pelvis has been remodelled to enable bipedal walking. Therefore, there has been a trade-off between prenatal brain development, i.e., the size of the fetal head, and the diameter of the birth canal [109]. As a result, the volume of the brain at birth is about a quarter of adult size compared to 40% in the chimpanzee [110]. Indeed, in humans, the fetal pattern of brain growth continues for a year after birth [94]. Consequently, the fontanelles separating the bones of the skull do not close until 18 months to 2 years after birth [111]. The development of most other organs is as complete at birth as in other primates, all of which deliver precocial young.
In contrast, rodents such as mouse and rat deliver truly altricial young with closed eyes, naked skin, and incomplete development of major organs, such as the kidneys [2]. The short gestation means that there is no period equivalent to the third trimester of human pregnancy when obstetric complications are most evident. Differences in gestation length are even reflected in placental function, the different role of placental lactogens in rodents and primates being a case in point.

7. Discussion

7.1. Placental Evolution

Placental characters shared with all eutherian mammals are invasive placentation and the decidual reaction (Table 2). Persistence of decidua into late gestation is common to the major clade Euarchontoglires, which includes rodents as well as primates. Many characteristics are shared with the primate suborder Haplorrhini, including features of the fetal membranes that Hubrecht used to justify classing tarsiers with monkeys and apes [11]. Characters shared with Old World monkeys include villous placentation with an intervillous space and some aspects of trophoblast invasion. However, implantation is superficial in all primates except gibbons and great apes, and trophoblast invasion by the interstitial route is shared only with the great apes. This leads to the evolutionary timeline shown in Table 2.

7.2. Pregnancy Complications

Comparatively little is known about pregnancy complications in nonhuman primates. Preeclampsia may occur in great apes, but the evidence is thin, although in one case supported by a renal biopsy [113]. There is, however, much to be said for the argument that deep trophoblast invasion, especially by the interstitial route, can be linked to the emergence of preeclampsia in the great apes [51].
Many changes reminiscent of preeclampsia could be replicated in a baboon model by uterine artery ligation [114], but these may merely reflect responses to reduced oxygen delivery. Hypertension can develop spontaneously in the vervet monkey, even when not pregnant [115], and gestational hypertension in the closely related patas monkey (Erythrocebus patas) was accompanied by preeclampsia-like symptoms [116].
Fetal growth restriction is another focus of obstetric research. It occurs in New World monkeys that regularly bear twins or triplets, and the effects on birth weight and neonatal outcomes are currently under investigation [117]. Indeed, the common marmoset (Callithrix jacchus) is a promising model for pregnancy research [2].

8. Conclusions

The many unique features of human pregnancy and placentation pose problems in planning and interpreting animal experiments. Two factors are involved. The first is phylogenetic distance. Quite a few features are shared with haplorrhine primates, making them the models of choice. Baboons and macaques share additional features such as endovascular trophoblast and spiral artery transformation as well as a true intervillous space. On the other hand, maintenance of breeding colonies is costly. Therefore, it is worth considering the common marmoset for which caging and feeding costs are much lower [118]. As mentioned in the introduction, primates and rodents last shared a common ancestor in the Cretaceous period, so it is not surprising that placental evolution has pursued different paths. There has, for example, been convergent evolution of placental lactogens to serve different purposes.
A second factor compounds the problem with rodent models. This is the difference in reproductive strategies. The short generation times of mice and rats make them ideal laboratory animals. Unfortunately, the same qualities render them unsatisfactory for pregnancy research [119,120]. The major obstetric syndromes become manifest in the third trimester, but there is no equivalent period in the mouse. The newborn mouse is truly altricial, with much of organ development occurring in the postnatal period. I have been at pains to stress, as argued by Martin [111], that human babies are precocial in almost all aspects save brain development; they are not altricial. Alternative rodent models are the spiny mouse (Acomys cahirinus) and guinea pig, both of which deliver precocial young [2].

Funding

This research received no external funding.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Brosens, I.; Pijnenborg, R.; Vercruysse, L.; Romero, R. The “Great Obstetrical Syndromes” are associated with disorders of deep placentation. Am. J. Obstet. Gynecol. 2011, 204, 193–201. [Google Scholar] [CrossRef] [Green Version]
  2. Carter, A.M. Animal models of human pregnancy and placentation: Alternatives to the mouse. Reproduction 2020, 160, R129–R143. [Google Scholar] [CrossRef]
  3. Carter, A.M. Animal models of human placentation—A review. Placenta 2007, 28, S41–S47. [Google Scholar] [CrossRef]
  4. Carter, A.M. Evolution of Placental Function in Mammals: The Molecular Basis of Gas and Nutrient Transfer, Hormone Secretion, and Immune Responses. Physiol. Rev. 2012, 92, 1543–1576. [Google Scholar] [CrossRef]
  5. Huhn, O.; Zhao, X.; Esposito, L.; Moffett, A.; Colucci, F.; Sharkey, A.M. How Do Uterine Natural Killer and Innate Lymphoid Cells Contribute to Successful Pregnancy? Front. Immunol. 2021, 12, 607669. [Google Scholar] [CrossRef] [PubMed]
  6. Prabhudas, M.; Bonney, E.; Caron, K.; Dey, S.; Erlebacher, A.; Fazleabas, A.; Fisher, S.; Golos, T.; Matzuk, M.; McCune, J.M.; et al. Immune mechanisms at the maternal-fetal interface: Perspectives and challenges. Nat. Immunol. 2015, 16, 328–334. [Google Scholar] [CrossRef] [PubMed]
  7. Meredith, R.W.; Janečka, J.E.; Gatesy, J.; Ryder, O.A.; Fisher, C.A.; Teeling, E.C.; Goodbla, A.; Eizirik, E.; Simão, T.L.L.; Stadler, T.; et al. Impacts of the Cretaceous Terrestrial Revolution and KPg Extinction on Mammal Diversification. Science 2011, 334, 521–524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Murphy, W.J.; Pringle, T.H.; Crider, T.A.; Springer, M.S.; Miller, W. Using genomic data to unravel the root of the placental mammal phylogeny. Genome Res. 2007, 17, 413–421. [Google Scholar] [CrossRef] [Green Version]
  9. Janečka, J.E.; Miller, W.; Pringle, T.H.; Wiens, F.; Zitzmann, A.; Helgen, K.M.; Springer, M.S.; Murphy, W.J. Molecular and Genomic Data Identify the Closest Living Relative of Primates. Science 2007, 318, 792–794. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Carter, A.M.; Enders, A.C. The Evolution of Epitheliochorial Placentation. Annu. Rev. Anim. Biosci. 2013, 1, 443–467. [Google Scholar] [CrossRef]
  11. Hubrecht, A.A.W. Relations of tarsius to the lemurs and apes. Science 1897, 5, 550–551. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Perelman, P.; Johnson, W.; Roos, C.; Seuánez, H.N.; Horvath, J.E.; Moreira, M.A.M.; Kessing, B.; Pontius, J.; Roelke, M.; Rumpler, Y.; et al. A Molecular Phylogeny of Living Primates. PLoS Genet. 2011, 7, e1001342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. James, J.; Carter, A.; Chamley, L. Human placentation from nidation to 5 weeks of gestation. Part I: What do we know about formative placental development following implantation? Placenta 2012, 33, 327–334. [Google Scholar] [CrossRef] [PubMed]
  14. Hill, J.P., II. Croonian lecture—The developmental history of the primates. Philos. Trans. R. Soc. Lond. Ser. B Contain. Pap. A Biol. Character 1932, 221, 45–178. [Google Scholar] [CrossRef]
  15. Selenka, E. Entwickelung des Gibbon (Hylobates und Siamanga). Stud. Über Enwickelungsgeschichte Tiere 1899, 8, 163–208. [Google Scholar]
  16. Carter, A.M.; Enders, A.C.; Pijnenborg, R. The role of invasive trophoblast in implantation and placentation of primates. Philos. Trans. R. Soc. B Biol. Sci. 2015, 370, 20140070. [Google Scholar] [CrossRef]
  17. Dupressoir, A.; Lavialle, C.; Heidmann, T. From ancestral infectious retroviruses to bona fide cellular genes: Role of the captured syncytins in placentation. Placenta 2012, 33, 663–671. [Google Scholar] [CrossRef]
  18. Lavialle, C.; Cornelis, G.; Dupressoir, A.; Esnault, C.; Heidmann, O.; Vernochet, C.; Heidmann, T. Paleovirology of ‘syncytins’, retroviral env genes exapted for a role in placentation. Philos. Trans. R. Soc. B Biol. Sci. 2013, 368, 20120507. [Google Scholar] [CrossRef] [Green Version]
  19. Chavan, A.R.; Bhullar, B.-A.S.; Wagner, G.P. What was the ancestral function of decidual stromal cells? A model for the evolution of eutherian pregnancy. Placenta 2016, 40, 40–51. [Google Scholar] [CrossRef] [Green Version]
  20. Mossman, H.W. Vertebrate Fetal Membranes: Comparative Ontogeny and Morphology; Evolution; Phylogenetic Significance; Basic Functions; Research Opportunities; Rutgers University Press: New Brunswick, NJ, USA, 1987. [Google Scholar]
  21. Kin, K.; Maziarz, J.; Chavan, A.R.; Kamat, M.; Vasudevan, S.; Birt, A.; Wagner, G.P. The transcriptomic evolution of mammalian pregnancy: Gene expression innovations in endometrial stromal fibroblasts. Genome Biol. Evol. 2016, 8, 2459–2473. [Google Scholar] [CrossRef] [Green Version]
  22. Griffith, O.W.; Chavan, A.R.; Pavlicev, M.; Protopapas, S.; Callahan, R.; Maziarz, J.; Wagner, G.P. Endometrial recognition of pregnancy occurs in the grey short-tailed opossum (Monodelphis domestica). Proc. Biol. Sci. 2019, 286, 20190691. [Google Scholar] [CrossRef]
  23. Griffith, O.W.; Chavan, A.R.; Protopapas, S.; Maziarz, J.; Romero, R.; Wagner, G.P. Embryo implantation evolved from an ancestral inflammatory attachment reaction. Proc. Natl. Acad. Sci. USA 2017, 114, E6566–E6575. [Google Scholar] [CrossRef] [Green Version]
  24. Erkenbrack, E.M.; Maziarz, J.D.; Griffith, O.; Liang, C.; Chavan, A.R.; Nnamani, M.C.; Wagner, G.P. The mammalian decidual cell evolved from a cellular stress response. PLoS Biol. 2018, 16, e2005594. [Google Scholar] [CrossRef]
  25. Enders, A.C.; King, B.F. Development of the human yolk sac. In The Human Yolk Sac and Yolk Sac Tumors; Nogales, F.F., Ed.; Springer: Berlin, Germany, 1993; pp. 33–47. [Google Scholar]
  26. Burton, G.J.; Cindrova-Davies, T.; Turco, M.Y. Review: Histotrophic nutrition and the placental-endometrial dialogue during human early pregnancy. Placenta 2020, 102, 21–26. [Google Scholar] [CrossRef]
  27. Boroviak, T.; Stirparo, G.; Dietmann, S.; Hernando-Herraez, I.; Mohammed, H.; Reik, W.; Smith, A.; Sasaki, E.; Nichols, J.; Bertone, P. Single cell transcriptome analysis of human, marmoset and mouse embryos reveals common and divergent features of preimplantation development. Development 2018, 145, 27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Ross, C.; Boroviak, T.E. Origin and function of the yolk sac in primate embryogenesis. Nat. Commun. 2020, 11, 1–14. [Google Scholar] [CrossRef]
  29. Carter, A. IFPA Senior Award Lecture: Mammalian fetal membranes. Placenta 2016, 48, S21–S30. [Google Scholar] [CrossRef] [PubMed]
  30. Mess, A.; Carter, A.M. Evolutionary transformations of fetal membrane characters in Eutheria with special reference to Afrotheria. J. Exp. Zool. Part B Mol. Dev. Evol. 2006, 306B, 140–163. [Google Scholar] [CrossRef] [PubMed]
  31. Bongso, T.A.; Basrur, P.K. Foetal fluids in cattle. Can. Vet. J. 1976, 17, 38–41. [Google Scholar] [PubMed]
  32. Wildman, D.E.; Chen, C.; Erez, O.; Grossman, L.I.; Goodman, M.; Romero, R. Evolution of the mammalian placenta revealed by phylogenetic analysis. Proc. Natl. Acad. Sci. USA 2006, 103, 3203–3208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Jones, C.; Jauniaux, E. Ultrastructure of the materno-embryonic interface in the first trimester of pregnancy. Micron 1995, 26, 145–173. [Google Scholar] [CrossRef]
  34. Enders, A.C. A comparative study of the fine structure of the trophoblast in several hemochorial placentas. Am. J. Anat. 1965, 116, 29–67. [Google Scholar] [CrossRef]
  35. Carter, A.M. Placental Gas Exchange and the Oxygen Supply to the Fetus. Compr. Physiol. 2015, 5, 1381–1403. [Google Scholar] [CrossRef] [PubMed]
  36. Mossman, H.W. The rabbit placenta and the problem of placental transmission. Am. J. Anat. 1926, 37, 433–497. [Google Scholar] [CrossRef]
  37. Metcalfe, J.; Bartels, H.; Moll, W. Gas exchange in the pregnant uterus. Physiol. Rev. 1967, 47, 782–838. [Google Scholar] [CrossRef]
  38. Adamson, S.L.; Lu, Y.; Whiteley, K.J.; Holmyard, D.; Hemberger, M.; Pfarrer, C.; Cross, J.C. Interactions between trophoblast cells and the maternal and fetal circulation in the mouse placenta. Dev. Biol. 2002, 250, 358–373. [Google Scholar] [CrossRef] [PubMed]
  39. Ramsey, E.M.; Harris, J.W.S. Comparison of utero-placental vasculature and circulation in the rhesus monkey and man. Contrib. Embryol. Carnegie Inst. 1966, 38, 61–70. [Google Scholar]
  40. Harris, L. IFPA Gabor Than Award lecture: Transformation of the spiral arteries in human pregnancy: Key events in the remodelling timeline. Placenta 2011, 32, S154–S158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Pijnenborg, R.; Vercruysse, L.; Hanssens, M. The Uterine Spiral Arteries in Human Pregnancy: Facts and Controversies. Placenta 2006, 27, 939–958. [Google Scholar] [CrossRef] [PubMed]
  42. Lash, G.E.; Bulmer, J.N. Do uterine natural killer (uNK) cells contribute to female reproductive disorders? J. Reprod. Immunol. 2011, 88, 156–164. [Google Scholar] [CrossRef]
  43. Burton, G.; Woods, A.; Jauniaux, E.; Kingdom, J. Rheological and Physiological Consequences of Conversion of the Maternal Spiral Arteries for Uteroplacental Blood Flow during Human Pregnancy. Placenta 2009, 30, 473–482. [Google Scholar] [CrossRef] [Green Version]
  44. Pankiewicz, K.; Fijałkowska, A.; Issat, T.; Maciejewski, T. Insight into the Key Points of Preeclampsia Pathophysiology: Uterine Artery Remodeling and the Role of MicroRNAs. Int. J. Mol. Sci. 2021, 22, 3132. [Google Scholar] [CrossRef] [PubMed]
  45. Zybina, T.G.; Frank, H.-G.; Biesterfeld, S.; Kaufmann, P. Genome multiplication of extravillous trophoblast cells in human placenta in the course of differentiation and invasion into endometrium and myometrium. II. Mech. Polyploidization Tsitologiya 2004, 46, 640–648. [Google Scholar]
  46. Pijnenborg, R.; Bland, J.M.; Robertson, W.B.; Dixon, G.; Brosens, I. The pattern of interstitial trophoblastic invasion of the myometrium in early human pregnancy. Placenta 1981, 2, 303–316. [Google Scholar] [CrossRef]
  47. Pijnenborg, R.; Vercruysse, L.; Carter, A.M. Deep trophoblast invasion and spiral artery remodelling in the placental bed of the lowland gorilla. Placenta 2011, 32, 586–591. [Google Scholar] [CrossRef] [PubMed]
  48. Pijnenborg, R.; Vercruysse, L.; Carter, A.M. Deep trophoblast invasion and spiral artery remodelling in the placental bed of the chimpanzee. Placenta 2011, 32, 400–408. [Google Scholar] [CrossRef] [PubMed]
  49. Pijnenborg, R.; D’Hooghe, T.; Vercruysse, L.; Bambra, C. Evaluation of trophoblast invasion in placental bed biopsies of the baboon, with immunohistochemical localisation of cytokeratin, fibronectin, and laminin. J. Med. Primatol. 1996, 25, 272–281. [Google Scholar] [CrossRef] [PubMed]
  50. Blankenship, T.N.; Enders, A.C.; King, B.F. Trophoblastic invasion and the development of uteroplacental arteries in the macaque: Immunohistochemical localization of cytokeratins, desmin, type IV collagen, laminin, and fibronectin. Cell Tissue Res. 1993, 272, 227–236. [Google Scholar] [CrossRef]
  51. Carter, A.M. Comparative studies of placentation and immunology in non-human primates suggest a scenario for the evolution of deep trophoblast invasion and an explanation for human pregnancy disorders. Reproduction 2011, 141, 391–396. [Google Scholar] [CrossRef] [Green Version]
  52. Verkeste, C.; Slangen, B.; Daemen, M.; Van Straaten, H.; Kohnen, G.; Kaufmann, P.; Peeters, L. The extent of trophoblast invasion in the preplacental vasculature of the guinea-pig. Placenta 1998, 19, 49–54. [Google Scholar] [CrossRef]
  53. Redline, R.W.; Lu, C.Y. Localization of fetal major histocompatibility complex antigens and maternal leukocytes in murine placenta. Implications for maternal-fetal immunological relationship. Lab. Investig. 1989, 61, 27–36. [Google Scholar] [PubMed]
  54. Hu, D.; Cross, J.C. Development and function of trophoblast giant cells in the rodent placenta. Int. J. Dev. Biol. 2010, 54, 341–354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Ain, R.; Canham, L.N.; Soares, M.J. Gestation stage-dependent intrauterine trophoblast cell invasion in the rat and mouse: Novel endocrine phenotype and regulation. Dev. Biol. 2003, 260, 176–190. [Google Scholar] [CrossRef] [Green Version]
  56. Geusens, N.; Hering, L.; Verlohren, S.; Luyten, C.; Drijkoningen, K.; Taube, M.; Vercruysse, L.; Hanssens, M.; Dechend, R.; Pijnenborg, R. Changes in endovascular trophoblast invasion and spiral artery remodelling at term in a transgenic preeclamptic rat model. Placenta 2010, 31, 320–326. [Google Scholar] [CrossRef] [PubMed]
  57. Medawar, P.B. Some immunological and endocrinological problems raised by the evolution of viviparity in vertebrates. Symp. Soc. Exp. Biol. 1953, 7, 320–338. [Google Scholar]
  58. Lash, G.E.; Robson, S.C.; Bulmer, J.N. Review: Functional role of uterine natural killer (uNK) cells in human early pregnancy decidua. Placenta 2010, 31, S87–S92. [Google Scholar] [CrossRef]
  59. Abi-Rached, L.; Kuhl, H.; Roos, C.; Ten Hallers, B.; Zhu, B.; Carbone, L.; Walter, L. A small, variable, and irregular killer cell Ig-like receptor locus accompanies the absence of MHC-C and MHC-G in gibbons. J. Immunol. 2010, 184, 1379–1391. [Google Scholar] [CrossRef] [Green Version]
  60. Golos, T.G.; Bondarenko, G.I.; Dambaeva, S.V.; Breburda, E.E.; Durning, M. On the role of placental Major Histocompatibility Complex and decidual leukocytes in implantation and pregnancy success using non-human primate models. Int. J. Dev. Biol. 2010, 54, 431–443. [Google Scholar] [CrossRef] [Green Version]
  61. Thomson, G.; Adams, E.J.; Parham, P. Evidence for an HLA-C -like locus in the orangutan Pongo pygmaeus. Immunogenetics 1999, 49, 865–871. [Google Scholar] [CrossRef]
  62. Aguilar, A.M.O.; Guethlein, L.A.; Adams, E.J.; Abi-Rached, L.; Moesta, A.; Parham, P. Coevolution of Killer Cell Ig-Like Receptors with HLA-C To Become the Major Variable Regulators of Human NK Cells. J. Immunol. 2010, 185, 4238–4251. [Google Scholar] [CrossRef] [Green Version]
  63. Penman, B.S.; Moffett, A.; Chazara, O.; Gupta, S.; Parham, P. Reproduction, infection and killer-cell immunoglobulin-like receptor haplotype evolution. Immunogenetics 2016, 68, 755–764. [Google Scholar] [CrossRef] [Green Version]
  64. Hiby, S.E.; Apps, R.; Sharkey, A.; Farrell, L.E.; Gardner, L.; Mulder, A.; Claas, F.H.; Walker, J.; Redman, C.C.; Morgan, L.; et al. Maternal activating KIRs protect against human reproductive failure mediated by fetal HLA-C2. J. Clin. Investig. 2010, 120, 4102–4110. [Google Scholar] [CrossRef]
  65. Hiby, S.E.; Walker, J.; O’Shaughnessy, K.M.; Redman, C.W.; Carrington, M.; Trowsdale, J.; Moffett, A. Combinations of Maternal KIR and Fetal HLA-C Genes Influence the Risk of Preeclampsia and Reproductive Success. J. Exp. Med. 2004, 200, 957–965. [Google Scholar] [CrossRef]
  66. Khakoo, S.; Rajalingam, R.; Shum, B.P.; Weidenbach, K.; Flodin, L.; Muir, D.G.; Canavez, F.; Cooper, S.L.; Valiante, N.M.; Lanier, L.L.; et al. Rapid Evolution of NK Cell Receptor Systems Demonstrated by Comparison of Chimpanzees and Humans. Immunity 2000, 12, 687–698. [Google Scholar] [CrossRef] [Green Version]
  67. Wroblewski, E.E.; Parham, P.; Guethlein, L.A. Two to Tango: Co-evolution of Hominid Natural Killer Cell Receptors and MHC. Front. Immunol. 2019, 10, 177. [Google Scholar] [CrossRef] [Green Version]
  68. Le Bouteiller, P. HLA-G in the human placenta: Expression and potential functions. Biochem. Soc. Trans. 2000, 28, 208–212. [Google Scholar] [CrossRef]
  69. Hunt, J.S.; Langat, D.L. HLA-G: A human pregnancy-related immunomodulator. Curr. Opin. Pharmacol. 2009, 9, 462–469. [Google Scholar] [CrossRef] [Green Version]
  70. Castro, M.J.; Morales, P.; Fernández-Soria, V.; Suarez, B.; Recio, M.J.; Alvarez, M.; Arnaiz-Villena, A. Allelic diversity at the primate MHC-G locus: Exon 3 bears stop codons in all Cercopithecinae sequences. Immunogenetics 1996, 43, 327–336. [Google Scholar] [CrossRef] [PubMed]
  71. Bondarenko, G.I.; Dambaeva, S.V.; Grendell, R.L.; Hughes, A.L.; Durning, M.; Garthwaite, M.A.; Golos, T.G. Characterization of cynomolgus and vervet monkey placental MHC class I expression: Diversity of the nonhuman primate AG locus. Immunogenetics 2009, 61, 431–442. [Google Scholar] [CrossRef] [Green Version]
  72. Croy, B.A.; Luross, J.A.; Guimond, M.J.; Hunt, J.S. Uterine natural killer cells: Insights into lineage relationships and functions from studies of pregnancies in mutant and transgenic mice. Nat. Immun. 1996, 15, 22–33. [Google Scholar]
  73. Burke, S.; Barrette, V.F.; Gravel, J.; Carter, A.L.I.; Hatta, K.; Zhang, J.; Chen, Z.; Leno-Durán, E.; Bianco, J.; Leonard, S.; et al. Uterine NK Cells, Spiral Artery Modification and the Regulation of Blood Pressure During Mouse Pregnancy. Am. J. Reprod. Immunol. 2010, 63, 472–481. [Google Scholar] [CrossRef]
  74. Croy, B.; Esadeg, S.; Chantakru, S.; Heuvel, M.V.D.; Paffaro, V.A.; He, H.; Black, G.P.; Ashkar, A.; Kiso, Y.; Zhang, J. Update on pathways regulating the activation of uterine Natural Killer cells, their interactions with decidual spiral arteries and homing of their precursors to the uterus. J. Reprod. Immunol. 2003, 59, 175–191. [Google Scholar] [CrossRef]
  75. Nguyen-Lefebvre, A.T.; Ajith, A.; Portik-Dobos, V.; Horuzsko, D.D.; Mulloy, L.L.; Horuzsko, A. Mouse models for studies of HLA-G functions in basic science and pre-clinical research. Hum. Immunol. 2016, 77, 711–719. [Google Scholar] [CrossRef] [PubMed]
  76. Napso, T.; Yong, H.E.J.; Lopez-Tello, J.; Sferruzzi-Perri, A. The Role of Placental Hormones in Mediating Maternal Adaptations to Support Pregnancy and Lactation. Front. Physiol. 2018, 9, 1091. [Google Scholar] [CrossRef] [PubMed]
  77. Maston, G.A.; Ruvolo, M. Chorionic gonadotropin has a recent origin within primates and an evolutionary history of selection. Mol. Biol. Evol. 2002, 19, 320–335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Chopineau, M.; Stewart, F.; Allen, W.R. Cloning and analysis of the cDNA encoding the horse and donkey luteinizing hormone beta-subunits. Gene 1995, 160, 253–256. [Google Scholar] [CrossRef]
  79. Chen, E.Y.; Liao, Y.-C.; Smith, D.H.; Barrera-Saldaña, H.A.; Gelinas, R.E.; Seeburg, P.H. The human growth hormone locus: Nucleotide sequence, biology, and evolution. Genomics 1989, 4, 479–497. [Google Scholar] [CrossRef]
  80. Frankenne, F.; Closset, J.; Gomez, F.; Scippo, M.L.; Smal, J.; Hennen, G. The Physiology of Growth Hormones (GHs) in Pregnant Women and Partial Characterization of the Placental GH Variant. J. Clin. Endocrinol. Metab. 1988, 66, 1171–1180. [Google Scholar] [CrossRef] [PubMed]
  81. de Mendoza, A.R.; Escobedo, D.E.; Dávila, I.M.; Saldaña, H. Expansion and divergence of the GH locus between spider monkey and chimpanzee. Gene 2004, 336, 185–193. [Google Scholar] [CrossRef]
  82. Wallis, O.C.; Wallis, M. Evolution of growth hormone in primates: The GH gene clusters of the New World monkeys marmoset (Callithrix jacchus) and white-fronted capuchin (Cebus albifrons). J. Mol. Evol. 2006, 63, 591–601. [Google Scholar] [CrossRef] [PubMed]
  83. Papper, Z.; Jameson, N.M.; Romero, R.; Weckle, A.L.; Mittal, P.; Benirschke, K.; Santolaya-Forgas, J.; Uddin, M.; Haig, D.; Goodman, M.; et al. Ancient origin of placental expression in the growth hormone genes of anthropoid primates. Proc. Natl. Acad. Sci. USA 2009, 106, 17083–17088. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Soares, M.J. The prolactin and growth hormone families: Pregnancy-specific hormones/cytokines at the maternal-fetal interface. Reprod. Biol. Endocrinol. 2004, 2, 51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Csapo, A. Progesterone block. Am. J. Anat. 1956, 98, 273–291. [Google Scholar] [CrossRef]
  86. Thorburn, G.D.; Challis, J.R.G.; Robinson, J.S. Endocrine Control of Parturition; Springer Science and Business Media LLC: Tokyo, Japan, 1977; pp. 653–732. [Google Scholar]
  87. Merlino, A.A.; Welsh, T.N.; Tan, H.; Yi, L.J.; Cannon, V.; Mercer, B.M.; Mesiano, S. Nuclear Progesterone Receptors in the Human Pregnancy Myometrium: Evidence that Parturition Involves Functional Progesterone Withdrawal Mediated by Increased Expression of Progesterone Receptor-A. J. Clin. Endocrinol. Metab. 2007, 92, 1927–1933. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Shynlova, O.; Nadeem, L.; Zhang, J.; Dunk, C.; Lye, S. Myometrial activation: Novel concepts underlying labor. Placenta 2020, 92, 28–36. [Google Scholar] [CrossRef]
  89. Marinić, M.; Lynch, V.J. Relaxed constraint and functional divergence of the progesterone receptor (PGR) in the human stem-lineage. PLoS Genet. 2020, 16, e1008666. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Chen, C.; Opazo, J.C.; Erez, O.; Uddin, M.; Santolaya-Forgas, J.; Goodman, M.; Grossman, L.I.; Romero, R.; Wildman, D.E. The human progesterone receptor shows evidence of adaptive evolution associated with its ability to act as a transcription factor. Mol. Phylogenet. Evol. 2008, 47, 637–649. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Blanks, A.; Brosens, J. Progesterone Action in the Myometrium and Decidua in Preterm Birth. Facts Views Vis. ObGyn 2012, 4, 188–194. [Google Scholar]
  92. Challis, J.R.; John Davies, I.; Benirschke, K.; Hendrickx, A.G.; Ryan, K.J. The concentrations of progesterone, estrone and estradiol-17 beta in the peripheral plasma of the rhesus monkey during the final third of gestation, and after the induction of abortion with PGF 2 alpha. Endocrinology 1974, 95, 547–553. [Google Scholar] [CrossRef]
  93. Mitchell, B.F.; Taggart, M.J. Are animal models relevant to key aspects of human parturition? Am. J. Physiol. Integr. Comp. Physiol. 2009, 297, R525–R545. [Google Scholar] [CrossRef] [Green Version]
  94. Martin, R.D. The evolution of human reproduction: A primatological perspective. Am. J. Phys. Anthr. 2007, 134, 59–84. [Google Scholar] [CrossRef]
  95. Roberts, M. Growth, development, and parental care in the western tarsier (Tarsius bancanus) in captivity: Evidence for a “slow” life-history and nonmonogamous mating system. Int. J. Primatol. 1994, 15, 1–28. [Google Scholar] [CrossRef]
  96. Abbott, D.H.; Barnett, D.K.; Colman, R.J.; Yamamoto, M.E.; Schultz-Darken, N.J. Aspects of common marmoset basic biology and life history important for biomedical research. Comp. Med. 2003, 53, 339–350. [Google Scholar]
  97. Tardif, S.D.; Smucny, D.A.; Abbott, D.H.; Mansfield, K.; Schultz-Darken, N.; Yamamoto, M.E. Reproduction in captive common marmosets (Callithrix jacchus). Comp. Med. 2003, 53, 364–368. [Google Scholar]
  98. Altmann, J.; Samuels, A. Costs of maternal care: Infant-carrying in baboons. Behav. Ecol. Sociobiol. 1992, 29, 391–398. [Google Scholar] [CrossRef]
  99. Rhine, R.J.; Norton, G.W.; Wynn, G.M.; Wynn, R.D. Weaning of free-ranging infant baboons (Papio cynocephalus) as indicated by one-zero and instantaneous sampling of feeding. Int. J. Primatol. 1985, 6, 491–499. [Google Scholar] [CrossRef]
  100. Gesquiere, L.R.; Altmann, J.; Archie, E.A.; Alberts, S.C. Interbirth intervals in wild baboons: Environmental predictors and hormonal correlates. Am. J. Phys. Anthr. 2018, 166, 107–126. [Google Scholar] [CrossRef] [PubMed]
  101. Lappan, S. Patterns of Infant Care in Wild Siamangs (Symphalangus syndactylus) in Southern Sumatra. In The Gibbons; Springer Science and Business Media LLC: Tokyo, Japan, 2009; pp. 327–345. [Google Scholar]
  102. McConkey, K. Bornean Orangutan (Pongo pygmaeus). In World Atlas of Great Apes and Their Conservation; Caldecott, J.O., Miles, L., Eds.; University of California Press, in Association with UNEP-WCMC: Berkeley, CA, USA, 2005; pp. 161–183. [Google Scholar]
  103. Tullner, W.W. Comparative Aspects of Primate Chorionic Gonadotropins. In Reproductive Biology of the Primates; Luckett, W.P., Ed.; Karger: Basel, Switzerland, 1974; pp. 235–257. [Google Scholar]
  104. Ferriss, S.; Robbins, M.M.; Williamson, E.A. Eastern gorilla (Gorilla beringei). In World Atlas of Great Apes and Their Conservation; Caldecott, J.O., Miles, L., Eds.; University of California Press, in Association with UNEP-WCMC: Berkeley, CA, USA, 2005; pp. 129–152. [Google Scholar]
  105. Canington, S.L. Gorilla beringei (Primates: Hominidae). Mamm. Species 2018, 967, 119–133. [Google Scholar] [CrossRef]
  106. Inskipp, T. Chimpanzee (Pan troglodytes). In World Atlas of Great Apes and their Conservation; Caldecott, J.O., Miles, L., Eds.; University of California Press, in Association with UNEP-WCMC: Berkeley, CA, USA, 2005; pp. 53–81. [Google Scholar]
  107. Jones, C.; Jones, C.A.; Knox Jones, J.; Wilson, D.E. Pan troglodytes. Mamm. Species 1996, 529, 1–9. [Google Scholar] [CrossRef] [Green Version]
  108. Martin, R.D.; MacLarnon, A. Gestation period, neonatal size and maternal investment in placental mammals. Nature 1985, 313, 220–223. [Google Scholar] [CrossRef]
  109. Trevathan, W. Primate pelvic anatomy and implications for birth. Philos. Trans. R. Soc. B Biol. Sci. 2015, 370, 20140065. [Google Scholar] [CrossRef] [PubMed]
  110. Sakai, T.; Hirata, S.; Fuwa, K.; Sugama, K.; Kusunoki, K.; Makishima, H.; Eguchi, T.; Yamada, S.; Ogihara, N.; Takeshita, H. Fetal brain development in chimpanzees versus humans. Curr. Biol. 2012, 22, R791–R792. [Google Scholar] [CrossRef] [Green Version]
  111. Martin, R.D. How We Do It: The Evolution and Future of Human Reproduction, xii; Basic Books: New York, NY, USA, 2013; p. 304. [Google Scholar]
  112. Chatterjee, H.J.; Ho, S.Y.W.; Barnes, I.; Groves, C. Estimating the phylogeny and divergence times of primates using a supermatrix approach. BMC Evol. Biol. 2009, 9, 259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Stout, C.; Lemmon, W. Glomerular capillary endothelial swelling in a pregnant chimpanzee. Am. J. Obstet. Gynecol. 1969, 105, 212–215. [Google Scholar] [CrossRef]
  114. Makris, A.; Thornton, C.; Thompson, J.; Thomson, S.; Martin, R.; Ogle, R.; Waugh, R.; McKenzie, P.; Kirwan, P.; Hennessy, A. Uteroplacental ischemia results in proteinuric hypertension and elevated sFLT-1. Kidney Int. 2007, 71, 977–984. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Rhoads, M.K.; Goleva, S.B.; Beierwaltes, W.H.; Osborn, J.L. Renal vascular and glomerular pathologies associated with spontaneous hypertension in the nonhuman primate Chlorocebus aethiops sabaeus. Am. J. Physiol. Integr. Comp. Physiol. 2017, 313, R211–R218. [Google Scholar] [CrossRef] [Green Version]
  116. Palmer, A.E.; London, W.T.; Sly, D.L.; Rice, J.M. Spontaneous preeclamptic toxemia of pregnancy in the patas monkey (Erythrocebus patas). Lab. Anim. Sci. 1979, 29, 102–106. [Google Scholar] [PubMed]
  117. Rutherford, J.N.; Tardif, S.D. Placental efficiency and intrauterine resource allocation strategies in the common marmoset pregnancy. Am. J. Phys. Anthr. 2008, 137, 60–68. [Google Scholar] [CrossRef]
  118. Mansfield, K. Marmoset models commonly used in biomedical research. Comp. Med. 2003, 53, 383–392. [Google Scholar]
  119. Schmidt, A.; Prieto, D.M.M.; Pastuschek, J.; Fröhlich, K.; Markert, U.R. Only humans have human placentas: Molecular differences between mice and humans. J. Reprod. Immunol. 2015, 108, 65–71. [Google Scholar] [CrossRef]
  120. Malassine, A.; Frendo, J.-L.; Evain-Brion, D. A comparison of placental development and endocrine functions between the human and mouse model. Hum. Reprod. Updat 2003, 9, 531–539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. The mammalian tree. (A) The four major clades of eutherians [8]. (B) The orders of Euarchontoglires [9]. Note the separation of Glires (including rodents) from Euarchonta (including primates). There are alternative interpretations of the root of the tree and the position of tree shrews. Reprinted with permission from [2] © 2021 Society for Reproduction and Fertility.
Figure 1. The mammalian tree. (A) The four major clades of eutherians [8]. (B) The orders of Euarchontoglires [9]. Note the separation of Glires (including rodents) from Euarchonta (including primates). There are alternative interpretations of the root of the tree and the position of tree shrews. Reprinted with permission from [2] © 2021 Society for Reproduction and Fertility.
Ijms 22 08099 g001
Figure 2. Classification of primates [12]. Strepsirrhines and haplorrhines are suborders, tarsiers are regarded as an infraorder, whilst the other clades shown are superfamilies. OW, Old World; NW, New World. Reprinted with permission from [2] © 2021 Society for Reproduction and Fertility.
Figure 2. Classification of primates [12]. Strepsirrhines and haplorrhines are suborders, tarsiers are regarded as an infraorder, whilst the other clades shown are superfamilies. OW, Old World; NW, New World. Reprinted with permission from [2] © 2021 Society for Reproduction and Fertility.
Ijms 22 08099 g002
Figure 3. Early differentiation of mesoderm in the human embryo. (a) Trophoblastic plate stage (Carnegie Stage 5a) showing the inner cell mass (icm). The cavity of the blastocyst is collapsed. The pad of trophoblast has different sized nuclei and both cellular and syncytial trophoblast (syn tr) (Carnegie Embryo #8020). Scale bar, 70 µm. (b) Early lacunar stage (Carnegie Stage 5b) Some maternal blood has leaked into the primary yolk sac (pys). Note the irregular shape of the lacunae on the right, which appear to be formed from expanding clefts (Carnegie Embryo #8004). Scale bar, 90 µm. (c) Lacunar stage (Carnegie Stage 5c). Note the anastomotic lacunae within the syncytiotrophoblast. In the area between the trophoblast and the already partially constricted primary yolk sac (pys), there are mesenchymal cells (extraembryonic mesoderm) (Carnegie Embryo #7699). Scale bar, 176 µm. (d) Predecessors of the primary villi (Carnegie Stage 6). The cytotrophoblast (cyt troph) is accumulating in the partitions between lacunae, initiating the formation of primary villi (Carnegie Embryo #9260). Scale bar = 70 µm. Reprinted with permission from Carter, Enders and Pijnenborg [16] © The Authors. Published by the Royal Society. All rights reserved.
Figure 3. Early differentiation of mesoderm in the human embryo. (a) Trophoblastic plate stage (Carnegie Stage 5a) showing the inner cell mass (icm). The cavity of the blastocyst is collapsed. The pad of trophoblast has different sized nuclei and both cellular and syncytial trophoblast (syn tr) (Carnegie Embryo #8020). Scale bar, 70 µm. (b) Early lacunar stage (Carnegie Stage 5b) Some maternal blood has leaked into the primary yolk sac (pys). Note the irregular shape of the lacunae on the right, which appear to be formed from expanding clefts (Carnegie Embryo #8004). Scale bar, 90 µm. (c) Lacunar stage (Carnegie Stage 5c). Note the anastomotic lacunae within the syncytiotrophoblast. In the area between the trophoblast and the already partially constricted primary yolk sac (pys), there are mesenchymal cells (extraembryonic mesoderm) (Carnegie Embryo #7699). Scale bar, 176 µm. (d) Predecessors of the primary villi (Carnegie Stage 6). The cytotrophoblast (cyt troph) is accumulating in the partitions between lacunae, initiating the formation of primary villi (Carnegie Embryo #9260). Scale bar = 70 µm. Reprinted with permission from Carter, Enders and Pijnenborg [16] © The Authors. Published by the Royal Society. All rights reserved.
Ijms 22 08099 g003
Figure 4. The interhaemal barrier of the human placenta is classified as haemomonochorial. The intervillous space is separated from blood in the fetal capillary by syncytiotrophoblast and fetal capillary endothelium with their basal membranes. A very thin layer of connective tissue cytoplasm is interposed between the two basal membranes. Courtesy of Dr. Allen C. Enders. Reproduced with permission from [35] Copyright © American Physiological Society.
Figure 4. The interhaemal barrier of the human placenta is classified as haemomonochorial. The intervillous space is separated from blood in the fetal capillary by syncytiotrophoblast and fetal capillary endothelium with their basal membranes. A very thin layer of connective tissue cytoplasm is interposed between the two basal membranes. Courtesy of Dr. Allen C. Enders. Reproduced with permission from [35] Copyright © American Physiological Society.
Ijms 22 08099 g004
Table 1. Precocity and parental care in selected primates. Apart from tarsier and marmoset, data are from field observations on free-living populations. Gestation lengths are approximate and based on few observations.
Table 1. Precocity and parental care in selected primates. Apart from tarsier and marmoset, data are from field observations on free-living populations. Gestation lengths are approximate and based on few observations.
CladeSpeciesCommon NameLength of GestationParental CareReferences
TarsiersCephalopacus bancanusWestern tarsier178 daysNutritional and social independence by 60 days[95]
New World monkeysCallithrix jacchusCommon marmoset143–144 daysIndependent movement by 3 weeks; weaning by 3 months[96,97]
Old World monkeysPapio cynocephalusYellow baboon178 ± 6 daysMilk supplemented early with plant foods; fully weaned after about a year; carried for 8 months[98,99,100]
Lesser apesSymphalangus syndactylusSiamang230–235 daysPartial weaning at 6 months; travel independently by 1 year[101]
Great apesPongo pygmaeusBornean orangutan275 daysPartial weaning by 11 months; fully independent at 7–10 years [102,103]
Gorilla beringeiEastern gorilla255 daysWeaning at 3–4 years[103,104,105]
Pan troglodytesChimpanzee196–260 daysWeaning at 10 months; dependent on mother for 5 years[103,106,107]
Table 2. Characteristics of human placentation and their estimated appearance during evolution. Branching points (Mya, million years ago) are estimates based on molecular data [7,112].
Table 2. Characteristics of human placentation and their estimated appearance during evolution. Branching points (Mya, million years ago) are estimates based on molecular data [7,112].
CharacterTaxonomic CladeBranching Point Geological Period or EpochComments
Invasive placentationEutheria98.5 MyaLate Cretaceous
Decidual reaction Eutheria98.5 MyaLate CretaceousAn inflammatory response in marsupials
Persistence of decidual stromal cellsEuarchontoglires (includes rodents and primates)91.8 MyaLate Cretaceous
Precocious extraembryonic mesodermHaplorrhini44.8 MyaMiddle Eocene
Secondary yolk sacHaplorrhini44.8 MyaMiddle Eocene
Allantoic stalkHaplorrhini44.8 MyaMiddle EoceneMany mammals have an allantoic sac
Haemomonochorial placentationHaplorrhini44.8 MyaMiddle Eocene
Syncytin-2 env geneHaplorrhini44.8 MyaMiddle Eocene
Chorionic gonadotropinHaplorrhini44.8 MyaMiddle Eocene
Placental lactogens and growth hormoneHaplorrhini44.8 MyaMiddle EoceneVary between primate lineages
Trophoblast invasion by intravascular routeOld World monkeys and apes29.8 MyaOligocene
Villous placentation with an intervillous spaceOld World monkeys and apes29.8 MyaOligoceneTrabecular placentation in tarsiers and NW monkeys
Interstitial implantationLesser and greater apes20.2 MyaEarly Miocene
Syncytin-1 env geneLesser and greater apes20.2 MyaEarly Miocene
Trophoblast invasion by interstitial routeGreat apes15.1 MyaMiddle Miocene
HLA-CGreat apes15.1 MyaMiddle Miocene
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Carter, A.M. Unique Aspects of Human Placentation. Int. J. Mol. Sci. 2021, 22, 8099. https://doi.org/10.3390/ijms22158099

AMA Style

Carter AM. Unique Aspects of Human Placentation. International Journal of Molecular Sciences. 2021; 22(15):8099. https://doi.org/10.3390/ijms22158099

Chicago/Turabian Style

Carter, Anthony M. 2021. "Unique Aspects of Human Placentation" International Journal of Molecular Sciences 22, no. 15: 8099. https://doi.org/10.3390/ijms22158099

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop