You are on page 1of 25

Official reprint from UpToDate®

www.uptodate.com
© 2021 UpToDate, Inc. and/or its affiliates. All Rights Reserved.

Immunology of the maternal-fetal interface


Author: Vikki M Abrahams, PhD
Section Editors: Rebecca Marsh, MD, Charles J Lockwood, MD, MHCM
Deputy Editor: Elizabeth TePas, MD, MS

All topics are updated as new evidence becomes available and our peer review process is complete.

Literature review current through: Aug 2021. | This topic last updated: Apr 06, 2021.

INTRODUCTION

In pregnant females, local adaptation of the maternal immune system allows for successful
coexistence between the mother and the semi-allograft that is the fetus/placenta expressing
both maternal (self) and paternal (nonself) genes [1-4]. Cytotoxic adaptive immune responses
are diminished, bypassed, or even abrogated, while regulatory adaptive immunity is enhanced
[5,6]. By contrast, innate (natural) immunity remains intact, serving two purposes: one, to
continue to provide host defense against infection, and two, to interact with fetal tissues to
promote successful placentation and pregnancy [4,7-10].

An introduction to the immunology of pregnancy and the maternal-fetal interface is presented


in this topic review. Basic immunologic concepts are reviewed separately. (See "An overview of
the innate immune system" and "The adaptive cellular immune response: T cells and
cytokines".)

IMMUNE DEFENSE MECHANISMS OF THE PLACENTA AND EXTRAPLACENTAL


MEMBRANES

The placenta and fetal membranes are directly exposed to maternal blood and tissues. Thus,
unique features of the cells that comprise this interface must underlie the remarkable ability of
the genetically distinct fetal tissue to inhabit the maternal host.

Many immunologists studying reproduction agree that maternal immunity is not solely
antagonistic to trophoblast tissue [11,12]. Indeed, a maternal immune presence in the decidua
is essential for successful implantation [4]. Evidence of the opposite hypothesis (ie, that
trophoblasts have offensive mechanisms for actively killing maternal lymphocytes) is lacking. It
is clear, however, that the placenta is normally protected from the killing functions of maternal
cells through a number of immune-protective mechanisms [13].

Fetal trophoblast cells — Fetal trophoblast cells are the specific cell layer that protects the
embryo from those components of the maternal immune system dedicated to destroying
foreign tissues. The inner cell mass and resultant embryo are secluded and protected beneath a
layer of trophoblastic cells throughout pregnancy ( figure 1).

Trophoblast cells are derived from the external trophectoderm layer of the blastocyst and
develop into the placenta. Precursor trophoblast cells choose one of three developmental
pathways (see "Placental development and physiology" and "The placental pathology report"):

● They may remain quiescent in the villi as a pool of cells for future needs (villous
cytotrophoblast cells).

● They may proliferate and migrate/invade into the decidua, ultimately forming the chorion
membrane (extravillous cytotrophoblast cells). Extravillous trophoblast cells may also
interact with maternal decidual immune cells and invade into the maternal spiral arteries,
replacing the endothelial layer (endovascular trophoblast).

● They may merge into the developing syncytiotrophoblast through a process of cell fusion to
create a multinucleate cell layer. While this creates a barrier for the controlled movement
and transport of molecules, the placenta is not an impermeable barrier. There is evidence
for the bidirectional movement of maternal and fetal cells [14]. This may be explained, in
part, by the presence of gaps in the syncytiotrophoblast, particularly later in gestation [15].

These subpopulations of trophoblasts are variably exposed to maternal hematopoietic


elements in the decidua (extravillous cytotrophoblast cells, chorion membrane
cytotrophoblasts, and endovascular trophoblasts) and in maternal blood either flowing over the
fetal surface of the placenta (syncytiotrophoblasts) or within the spiral arteries (endovascular
trophoblasts). Certain trophoblast cells (villous cytotrophoblast cells) are only rarely exposed to
maternal blood.

Placental membranes — The placenta and its attached membranes can protect themselves
from maternal immune cell attack and modulate the local maternal immune system to promote
and maintain normal placentation and pregnancy. The mechanisms utilized vary with the state
of differentiation and anatomic location of the trophoblast cells. This cell type, which protects
the embryo itself as well as certain components of the extraembryonic membranes that are
derived from the inner cell mass, such as the amnion membrane, relies upon multiple
strategies to avoid detection by potentially cytotoxic maternal immune cells and antibody-
mediated cell destruction while influencing the local immune environment that is necessary for
successful pregnancy [4].

Specific immune-protective mechanisms

Altered expression of HLA and related molecules — The primary cellular immune


response that develops against transplanted tissue is directed against the major
histocompatibility complex (MHC) proteins on donor tissue [16]. In humans, the MHC proteins
are called human leukocyte antigens (HLAs).

The placenta is not a typical transplant, since proteins derived from HLA genes are not
expressed codominantly on trophoblast cell membranes, unlike other cell types. (See "Human
leukocyte antigens (HLA): A roadmap" and "Transplantation immunobiology".)

● HLA class I – Strict regulation of the expression of specific HLA class I molecules in
subpopulations of trophoblasts is believed to protect the fetus against maternal immune
cells programmed to attack cells expressing foreign (paternal) HLA class I antigens. The
extravillous trophoblasts migrating into the decidua lack expression of HLA-A or HLA-B
class Ia antigens that are primary stimulators of classical graft rejection and instead display
a unique pattern of class Ia HLA-C and the nonclassical HLA class Ib molecules, HLA-E, HLA-
F, and HLA–G [1,17,18]. The genes encoding HLA-E, HLA-F, and HLA–G antigens have few
alleles in comparison with HLA-A and HLA-B. HLA-G, HLA-C, and HLA-F are expressed by first
trimester extravillous trophoblast, and, as gestation proceeds, their expression weakens
and becomes intracellular. HLA-E is expressed by the extravillous trophoblast only in the
first trimester [17].

HLA-C, HLA-E, HLA-G, and HLA-F may act to dampen or modulate maternal immune
responses by interacting with killer immunoglobulin-like receptors (KIRs) on decidual NK
(dNK) cells, macrophages, and a subset of T cells and with the T cell receptor on CD8+ T
cells [1,18-21]. The consequences of these interactions include activation of pathways in NK
cells and macrophages that interfere with the killer functions of these cells [22,23]. At the
same time, HLA-E and HLA-G may activate pathways in dNK cells, macrophages, and T cells
that promote trophoblast migration and placentation [18,24].

One mechanism by which HLA-G influences dNK cell and T cell function is through a
process of trogocytosis, in which a lymphocyte transfers plasma membrane fragments
containing cell surface molecules from an antigen-presenting cell to its own surface via the
immunologic synapse [18]. These acquired HLA-G+ NK and T cells are immunosuppressive
[25]. The HLA class Ia antigen, HLA-C, is expressed on extravillous trophoblast and may
have important interactions with leukocytes in some mothers [1,26]. As an example, certain
combinations of maternal KIRs and fetal HLA-C variants correlate with low birth weight and
preeclampsia [27,28]. HLA-C expression may be regulated by Nod-like receptor family pyrin
domain-containing 2 (NLRP2) [29].

HLA-G mRNA can be alternatively spliced to yield membrane-bound and soluble variants.
Soluble HLA-G is associated with immunomodulatory functions and promotion of
placentation [18,30]. Antigen-presenting cells expressing either a soluble or membrane-
bound form of HLA-G repress T cell alloproliferation via this pathway [31]. Soluble HLA-G
(HLA-G5) has been identified in maternal serum, is biochemically unique among the HLA-G
isoforms, and is associated with specific immune modulation and the promotion of a pro-
placentation environment [32-36].

In contrast to migrating extravillous cells, syncytiotrophoblast forming the outermost layer


of the placental villi that is exposed to maternal blood does not express membrane-bound
HLA class I antigens [37]. Syncytiotrophoblast in term placentas lacks HLA class IA mRNA
(HLA-A, HLA-B, and HLA-C) and membrane-bound protein [1,18-20]. However, there is
evidence for HLA class IB (HLA-E, HLA-F, and HLA-G) mRNA and antigens in this continuous
cell layer both early and late in pregnancy [1,18,33,38,39].

● HLA class II – The genes that encode potentially dangerous paternally derived foreign HLA
class II HLA-D region molecules are entirely repressed in trophoblast cells. None of the
trophoblast subpopulations express HLA class II antigens either in vivo or in vitro. Control
over transcription of class II genes may be exerted by silencing of expression of the class II
transactivator (CIITA), a transacting factor that is essential for constitutive and interferon
(IFN) gamma-inducible MHC class II gene transcription [40].

● B7 family – Members of the B7 family of costimulatory molecules that have both


lymphocyte stimulatory and inhibitory functions are expressed on trophoblast cells in
human placentas and may play a critical role in maintaining tolerance to the fetus. The
B7H1 protein (programmed death ligand 1 [PD-L1]; CD274), which has lymphocyte
inhibitory properties, is expressed on syncytiotrophoblast and, therefore, positioned to
interfere with activation of lymphocytes circulating in maternal blood. Interactions between
villous and extravillous trophoblast PD-L1 (B7H1) and programmed cell death protein 1 (PD-
1; CD279) expressed by maternal lymphocytes promote T regulatory (Treg) cell
development and function and inhibit activated T cells (T helper cell type 17 [Th17] cells)
[41-43].

● IDO – Trophoblast cells produce indoleamine 2,3-dioxygenase (IDO), which depletes


tryptophan by promoting its catabolism. This is believed to inactivate T cells since they
require tryptophan. Although IDO is clearly important to survival of the mouse embryo,
studies in IDO-/- knockout mice have failed to reveal any impact on pregnancy [44].
Evidence is beginning to emerge in support of this potential mechanism in humans [45].

● TNF superfamily – Apoptosis-inducing members of the TNF supergene family may also
have important roles in protecting the placenta and its membranes by inducing apoptosis
in potentially cytotoxic T cells. The ligands identified in and/or on human trophoblasts
include TNF-alpha, FasL, and TRAIL, as well as a number of less well-characterized TNF
superfamily members [46-49]. Some of these latter ligands, which include B cell-activating
factor (BAFF), may positively influence pregnancy host defenses by supporting maternal
and/or fetal antibody production [50].

All of these molecules, which are expressed as both membrane and soluble forms, can kill
activated immune cells targeting the trophoblast by transducing an apoptotic signal via
specific receptors on activated leukocytes. FasL may be of particular importance since FasL
prevents immune cell attack by interacting with leukocyte receptors (eg, Fas) in other
organs such as the eye and the testis.

● Microparticles – Placental-derived microparticles, such as microvesicles and exosomes that


contain an array of placental proteins, mRNA, and microRNAs, are thought to play a role in
regulating the maternal immune system during pregnancy. Studies have shown that the
trophoblast secretes active FasL via exosomes and microvesicles [51,52]. Some B7 family
members, as well as HLA-G, can also be released from the placenta via exosomes [53]. The
release and content of these microparticles are increased and/or altered under certain
pathologic conditions, and, as such, they may be involved in the pathogenesis of pregnancy
complications such as preeclampsia [54-57]. Exosome-mediated transfer of placental-
specific microRNAs regulates trophoblast and maternal cell immunity to viral infections
[58]. A novel mechanism by which maternal immune cells can signal to the placental unit
via exosomes has been proposed [59].

Soluble immunomodulators — Maternal immune modulation during pregnancy may also


be conferred by the synthesis of immunosuppressive or immunoregulatory molecules. As
examples, the human placenta produces progesterone, prostaglandin E2 (PGE2), and
antiinflammatory cytokines such as interleukin (IL) 10 and IL-4 [1,19,20,60,61]. Progesterone
may drive placental production of these antiinflammatory cytokines in placental cells just as it
does in lymphocytes. For example, IL-10 appears to stimulate production of HLA-G [62], and
dysfunction of this pathway may be important in preeclampsia [63]. Studies using IL-10-/-
knockout mice have highlighted the potentially important protective role for this cytokine in
pregnancy [64]. Soluble thymic stromal lymphopoietin (TSLP) may also play a role. TSLP
secreted by trophoblast cells stimulates decidual dendritic cells (dDCs) to produce IL-10 and
chemokine (C-C motif) ligand 17 (CCL17) [65]. These activated dDCs induce T helper cell type 2
(Th2) differentiation of decidual T cells. Lower levels of TSLP are seen in miscarriage than in
normal pregnancy.
Complement proteins — Trophoblast cells express high levels of the following complement
regulatory proteins: CD46 (membrane cofactor protein [MCP]), CD55 (decay accelerating factor
[DAF]), and CD59 (membrane inhibitor of reactive lysis [MIRL]) [66,67]. Complement regulatory
proteins are critically important for protecting the extraembryonic tissues from maternal
antipaternal cytotoxic antibodies since complement activation leads to opsonization and
destruction of the immunologic target (the fetal cells). (See "Regulators and receptors of the
complement system".)

Mothers routinely produce high titers of antibodies to paternally derived HLA and unique
trophoblast antigens, such as the unique placental isoform of alkaline phosphatase. Antibody
induction of complement-mediated lysis is prevented by high expression of CD46 and DAF in
various subpopulations of trophoblast cells.

Evidence supporting a role for complement proteins in placental immune protection is provided
by genetic deletion experiments. In mice, the absence of the complement regulatory protein
derived from Crry gene leads to placental disruption due to complement deposition in the
placenta and lysis of cells at the implantation site, followed by a massive inflammatory reaction
and fetal demise [68].

LOCAL IMMUNE ADAPTATION IN THE PREGNANT UTERUS

Dramatic changes take place in the uterus during pregnancy that also help contribute to
immune acceptance of and/or crosstalk with the fetus/placenta to promote a successful
pregnancy.

Differences in lymphocyte populations — The first histologically apparent maternal


immunologic adjustment to the embryo is a dramatic change in the relative proportions of
leukocyte subpopulations in the uterus. The endometrial natural killer (NK) cell population shifts
from uterine NK (uNK) cells to decidual NK (dNK) cells (CD56brightCD16-), and this phenotype is
distinct from both CD56bright and CD56dim peripheral blood NK (pNK) cells [69]. The origin of
dNK cells remains controversial as it is unclear whether these cells are recruited from the
periphery, if the local environment drives their unique phenotype, or both. Furthermore, their
function is thought to be distinct given that many genes, including some with
immunomodulatory properties, are overexpressed in dNK cells when compared with their
peripheral counterparts [69]. Invading fetal trophoblasts become admixed with dNK cells,
macrophages, and dendritic cells that account for approximately 70, 20, and 2 percent,
respectively, of all cells in the decidua [70-72]. In addition, the uterine T cell population during
pregnancy expands across gestation and is mostly regulatory in nature [73].

Decidual macrophages exhibit phenotypic elasticity, adapting to the local microenvironment.


During the peri-implantation period, decidual macrophages display an M1 (inflammatory)
phenotype. During the period of placentation, there is a mixture of M1 (inflammatory) and M2
(antiinflammatory) decidual macrophages, which shift to predominantly M2 after placentation
[74,75]. The placenta may promote this via secreted factors [76]. At parturition, there is an
accumulation of M1 decidual macrophages [74,75]. There are also subsets outside of
conventional phenotyping [74,77]. While decidual macrophages may help to prevent uterine
infections in pregnant females, some studies suggest that trophoblast-macrophage crosstalk is
more important for promoting normal placentation by being involved in implantation, placental
development, immunoregulation, vascular remodeling, and tissue homeostasis [9,76,78-80].
Aberrant macrophage numbers and activation may play a role in pregnancy complications, such
as preeclampsia, intrauterine growth restriction (IUGR), or preterm birth [9,81,82].
Macrophages are also present throughout gestation in the placental villi and are known as
Hofbauer cells. These fetally derived immune cells also display a M2 phenotype, although they
can generate strong inflammatory responses to infectious triggers. Hofbauer cell numbers are
also altered in pregnancy complications such as preeclampsia and chorioamnionitis [83-85].

Similarly, the major roles of dNK cells may also be unique to pregnancy [86]. Although dNK cells
express a high level of cytotoxic factors, they are unable to form cytotoxic synapses to deliver
granule contents [87] and instead appear to play a role in trophoblast attraction and invasion,
decidual and placental angiogenesis and possibly fetal vasculogenesis, and vascular
modifications in the uterus [88-91]. A population of uNK cells that are found in repeated
pregnancies has been identified. These "memory-like" uNK cells may be involved in promoting
placentation [92]. One study proposed that uNK cells maintain endometrial homeostasis by
clearing senescent decidual cells [93]. As with macrophages, alterations in dNK cell numbers
and activation status may play a role in pregnancy complications, such as immunologic
infertility, recurrent spontaneous abortion, and preeclampsia [27,28,94-98]. Excessive dNK cell
recruitment and/or expansion, as well as elevated cytotoxic activity, have been associated with
pregnancy disorders such as implantation failure and miscarriage [97]. However, some studies
suggest the opposite that deficiencies in uNK numbers are associated with recurrent pregnancy
loss, highlighting the need to reassess approaches to NK cell profiling in patients with adverse
outcomes [99,100]. NK cell depletion compromises pregnancy in mice, further underscoring
their importance during pregnancy [101].

Innate lymphoid cells (ILCs) distinct from NK cells have been characterized in the human
decidua. During early pregnancy, group 3 ILCs (ILC3) are found. While their function is not fully
understood, ILC3s may establish functional interactions with stromal cells, regulate the
recruitment and function of other maternal immune cells, and contribute to vascular
remodeling [102,103]. At term, both ILC2s and ILC3s have been found in the decidua, and levels
may increase in preterm birth [104].

The role of decidual dendritic cells (dDCs) is less clear, although mouse studies have
demonstrated that these cells are critical for successful implantation and may also be involved
in remodeling of the maternal vasculature [6,105,106]. However, alternative views exist. One is
that dendritic cells may promote systemic immune tolerance during pregnancy [107]. Another
is that dendritic cells are trapped in the decidua to prevent the exposure of peripheral T cells to
fetal antigens [108]. Uterine dendritic cells are also thought to contribute to pregnancy success
by influencing NK cell function and the cytokine profile at the maternal-fetal interface [109].
Depletion of uterine dendritic cells in mice results in a severe impairment of implantation and
embryo resorptions, highlighting their importance [106]. Preeclampsia is also associated with
persistence of decidual basalis macrophages and recruitment of dendritic cells [110,111]. One
hindrance of our understanding of the presence and function of dendritic cells at the maternal-
fetal interface is that there are many subpopulations, and there is a lack of comprehensive
phenotyping. However, new immunophenotyping approaches are emerging [112].

Gamma-delta T cells and a population of double-negative T lymphocytes (CD4-/CD8-) have been


reported in pregnant uteri [113,114]. Their roles are unclear, although immunosuppressive
gamma-delta T cells may play a role in regulating the maternal immune system to protect the
maternal-fetal interface from aggressive immune responses [115,116]. Increased numbers of
gamma-delta T cells may be associated with pregnancy loss [116]. It was once thought that
CD8+ effector T cells in normal pregnancy were eliminated from or prevented from entering the
maternal-fetal interface. However, studies have suggested that there is a population of highly
differentiated CD8+ effector memory T cells within the pregnant decidua. The antigenic target
of these cells is still unclear, as is their function and whether they are protective [117-119].

CD4+CD25+ regulatory T cells (Tregs) are also present in the decidua of normal pregnant
females, and their presence and expansion during pregnancy are thought to be triggered in
both alloantigen-dependent and alloantigen–independent manners [5,120,121]. In human
cocultures of HLA-G+ extravillous trophoblast cells and CD4+ T cells, there is an increase in the
number of cells expressing a Treg phenotype [122]. These fetal-specific Tregs persist after
delivery and rapidly reaccumulate during subsequent pregnancies, indicating a memory
response [123]. In a mouse model, increased numbers of Tregs are seen in response to an
antigenic fetus [124]. Selective killing of these Tregs results in fewer antigenic offspring. There
are decreased numbers of peripheral blood and decidual Treg cells in females with
preeclampsia compared with normal pregnancy subjects. These findings suggest that Tregs
play a role in maternal tolerance to the fetus [121,125,126].

A subpopulation of CD4+ interleukin (IL) 17-producing T cells (T helper cell type 17 [Th17]) have
also been described in pregnancy. Their numbers are also expanded in the pregnant uterus,
although not as much as CD4+CD25+ Tregs. While they are inflammatory in nature, the presence
of Th17 cells may play a role in protecting the maternal-fetal interface from microbes, and the
Tregs present may serve to regulate their function. Thus, altered numbers of Th17 and/or ratio
of Th17 to Tregs are associated with pregnancy complications, such a spontaneous abortion,
preeclampsia, and preterm birth [121,127].

Mast cells are classically associated with allergic immune responses. However, in pregnancy,
their presence in the decidua may contribute to successful placentation. Uterine mast cells
expand in early pregnancy in the mouse [128] and are higher in human pregnancies compared
with nonpregnant females [129]. In the mouse, they have been shown to promote an
antiinflammatory state and contribute to tissue remodeling, angiogenesis, and spiral artery
transformation [128].

A subset of IL-10-producing regulatory B cells (CD19+CD5+CD1d+) has been shown to expand


peripherally in mice and humans during pregnancy. In abortion-prone mice, expansion of these
regulatory B10 cells is not seen, and fetal rejection can be prevented by adoptive transfer of
these cells from normal pregnant mice [130,131]. While the presence of CD19+ B cells in the
human decidua throughout pregnancy has been reported [132,133], the importance of
regulatory B cells in the human decidua remains to be determined.

Soluble immunomodulatory agents — Uterine immune regulation is also provided by the


induction of immunomodulatory molecules that permeate the uterine environment. These
principally include progesterone, prostaglandins, and some cytokines.

Progesterone — Progesterone, the dominant hormone of pregnancy, is initially produced by


the corpus luteum. Subsequently, the placenta is responsible for almost all progesterone
synthesis.

High concentrations of progesterone can suppress the maternal immune response [134]. As an
example, progesterone alters the T helper cell type 1 (Th1)/T helper cell type 2 (Th2) balance
and inhibits production of tumor necrosis factor (TNF) alpha in both mouse and human
macrophages [135,136].
Prostaglandin E2 — Prostaglandin E2 (PGE2) is produced by resident macrophages and
decidual cells. Lymphocytes proliferate poorly in the presence of this compound.

Cytokines — High levels of Th2-type cytokines are typical of mouse pregnancy [137,138] but
are less definitive in pregnant females. Nonetheless, many still consider human pregnancy to
be a Th2 antiinflammatory condition and that a shift towards Th1 cytokines will lead to abortion
or pregnancy complications. As an example, elevated levels of IL-6 in cervicovaginal and
amniotic fluid, but not plasma, was associated with spontaneous preterm birth [139]. Similarly
high levels of proinflammatory IL-1-beta and TNF-alpha in the amniotic fluids are associated
with preterm birth [140,141]. While disruption in cytokine profiles during pregnancy may be
detrimental, it is important to appreciate that human pregnancy is both proinflammatory and
antiinflammatory, depending upon the stage of gestation, rather than focusing on the murine
Th1/Th2 terminology [4,120,142]. What is clear is that the appropriate balance of cytokine and
chemokine expression at the maternal-fetal interface can govern the immune cell profile and
function within the decidua. For example, one study demonstrated in mice that effector T cells
cannot accumulate within the decidua, in part because of the epigenetic silencing of key
chemokine genes in decidual stromal cells [143]. In contrast, the expression of specific
cytokines and chemokines by the decidua and placenta are critical for recruitment and
maintenance of pro-pregnancy immune cells [144].

MATERNAL SYSTEMIC IMMUNE RESPONSES

The pregnant immune system bears markers of both immune activation and dampening.
However, the consensus is that there is no generalized immunosuppression of maternal
immune responses in pregnancy. However, selective suppression or modulation may occur.
Studies have reported a decrease in proinflammatory capacity with dampening of the response
to microbial stimulation, while others have reported more monocyte-derived interleukin (IL) 12
in response to lipopolysaccharide (LPS) [145-149], for example. Thus, it appears that skewing
adjusts the maternal response to microbial challenge rather than effecting a global
suppression. In addition, the release of placental-derived microparticles may play a key role in
systemic maternal immune regulation [54-56].

Although multiparous females are excellent sources of antibodies to paternal human leukocyte
antigens (HLAs), maternal B lymphocytes specific for paternal HLA are partially deleted during
pregnancy. In addition, T lymphocytes specific for paternal HLA are difficult to demonstrate.
Results in transgenic mice suggest that pregnancy selectively depresses maternal T cells that
recognize paternal H-2 class I histocompatibility antigen [150]. (See "The adaptive humoral
immune response".)

There are at least two mechanisms for overcoming most immune reactions: One is active
suppression, and the other is enhanced tolerance. Enhanced tolerance has been clearly
demonstrated in normal pregnancy. Regulatory T cells (Tregs), which are critical mediators of
tolerance, become more numerous in pregnancy in response to the introduction of fetal
(paternal) antigens [151-154]. These Tregs produce IL-10, which appears to play a role in
maintaining pregnancy. In animals, IL-10 blockade increased the rate of abortions, although
successful pregnancy was still possible [155,156]. A population of IL-10 producing CD19+CD24hi
CD27+ regulatory B cells expands during normal pregnancy, and their role may be to suppress
undesired immune responses from maternal T cells [157].
IMMUNE FACTORS IN EARLY PREGNANCY LOSS

A high number of potential pregnancies are lost prior to or during implantation, based upon
the identification of chorionic gonadotropin (human growth hormone [hCG]) in females.
Spontaneous miscarriage before 20 weeks of gestation affects 15 to 20 percent of pregnant
females in the US [158]. While approximately 60 percent of miscarriages can be accounted for
by chromosomally abnormal embryos, the remainder are idiopathic [158] and, in many cases,
are thought to arise from implantation failure [159]. Whether any of these losses can be
accurately attributed to an immunologic cause (ie, based on maternal recognition and rejection
of normal embryos expressing paternal antigens) remains unclear despite research. In contrast,
recurrent embryonic losses due to genetic and endocrine abnormalities are well documented,
as are miscarriages due to infection and antiphospholipid antibodies. (See "Pregnancy loss
(miscarriage): Terminology, risk factors, and etiology" and "Causes of female infertility".)

Implantation and maternal immune rejection — Programming the uterus for successful


implantation of the semiallogeneic pregnancy may occur with the introduction of semen, which
contains not only paternal antigen but immunomodulatory factors, such as prostaglandin E2
(PGE2) and transforming growth factor (TGF) beta [160]. Subsequently, assuming that the
implanted blastocyst is intact and fully competent to develop into a fetus, the embryo should be
entirely protected by trophoblasts via the proposed mechanisms previously described.
However, in some females in whom the blastocyst is genetically abnormal, exposed paternally
derived antigens may be detected, resulting in a graft rejection response. A secondary immune
response would be expected to cause early rejection in cases of recurrent spontaneous
abortion.

Alternatively, some mothers with recurrent spontaneous abortions may lack essential
components of the described networks that provide immunologic protection to the embryos.

The most well-described immunologic dysregulation in females with infertility and recurrent
pregnancy losses (RPL) is related to natural killer (NK) cell numbers and function. In general,
excessive uterine NK (uNK) cell recruitment and/or expansion, as well as elevated cytotoxic
activity, are associated with implantation failure and RPL. Similarly, altered peripheral blood NK
(pNK) cells populations may also correlate with pregnancy losses and are a potential diagnostic
marker [97]. Consequently, there have been efforts targeting NK cells in an attempt to cure
infertility and miscarriage; however, well-controlled trials are lacking [161,162]. Alterations of
the T helper cell type 17 (Th17)/regulatory T cell (Treg) ratio both locally at the maternal-fetal
interface and systemically are also associated with RPL [97], as are elevated levels of peripheral
blood myeloid dendritic cells [163].

In addition to altered immune cell populations, disrupted immunoregulatory mechanisms are


associated with infertility and pregnancy loss. Certain combinations of maternal killer
immunoglobulin receptors (KIRs) and fetal HLA-C variants correlate with pregnancy loss [164],
as does reduced HLA-G expression [18,165]. Studies in mice implicate the deposition of
complement C3b on the placental vasculature early in gestation in conceptus demise [166], and,
in humans, elevated placental complement activity is linked to pregnancy loss [167]. Alterations
in expression of B7 costimulatory pathways and apoptosis-inducing tumor necrosis factor (TNF)
superfamily members may also be disrupted in pregnancy losses [168,169].

Infection — Interruption of pregnancy by a variety of mechanisms is most common during the


periimplantation period. Infection of the decidua, placenta, and fetal membranes appears to be
one cause of early pregnancy loss and preterm birth [170-172]. Infection at the maternal-fetal
interface may directly activate the trophoblast, decidual stroma, or chorioamniotic membranes
to generate either a proinflammatory or proapoptotic response that in turn may lead to a
disruption in the normal distribution, phenotype, and function of the decidual immune cells.
Similarly, infection may impact the function of gestational tissues and cell types. The pathways
leading from infection to preterm labor are not completely understood, although there is
increasing evidence to suggest that innate immune receptors, such as Toll-like receptors and
Nod-like receptors, may play a pivotal mechanistic role [4,173,174]. (See "Spontaneous preterm
birth: Pathogenesis".)

Autoimmune disease and recurrent pregnancy loss — Females with autoimmune disease


are at high risk for adverse pregnancy outcomes, such as recurrent pregnancy loss, as well as
late gestational complications, such as preeclampsia. (See "Pregnancy in women with systemic
lupus erythematosus" and "Thrombocytopenia in pregnancy" and "Hypothyroidism during
pregnancy: Clinical manifestations, diagnosis, and treatment" and "Management of myasthenia
gravis in pregnancy" and "Antiphospholipid syndrome: Pregnancy implications and
management in pregnant women".)

In particular, females with antiphospholipid syndrome (APS) and antiphospholipid antibodies


(aPL) are at high risk for recurrent spontaneous miscarriage, preeclampsia, placental
insufficiency, and intrauterine growth restriction (IUGR) [175]. In vivo animal studies have
shown that aPL target the trophoblast and decidua and trigger elevated local and systemic TNF-
alpha levels, elevated tissue factor expression and complement C3 deposition within the
decidua, and a neutrophilic infiltration in the decidua [176-181]. Blocking TNF-alpha, the
complement pathway, or tissue factor prevents aPL-mediated pregnancy failure and the
associated inflammation [176,178,179], suggesting that pregnancy failure in APS patients is a
result of inflammation rather than thrombosis at the maternal-fetal interface. An ongoing
clinical trial is testing anti-TNF-alpha therapy, certolizumab (IMPACT; NCT03152058) in females
with obstetric APS [182]. There is also a case report of the use of the complement inhibitor,
eculizumab, in a pregnant patient with APS [183]. Studies using human first trimester
trophoblast have demonstrated that aPL trigger an inflammatory response via the Toll-like
receptor 4 (TLR4) and NLRP3 inflammasome signaling pathways. In addition, aPL inhibit the
ability of the trophoblast to migrate and alter the cell's angiogenic factor production [184-187].
Thus, aPL may directly alter trophoblast function, induce placental inflammation, and alter the
immune cell profile at the maternal-fetal interface through innate immune pathways.

Recurrent pregnancy loss and therapeutic approaches — Immunization of potential


mothers with paternal or third-party leukocytes to improve reproductive success was, in the
past, a popular method of putatively improving livebirth rates [188]. This strategy developed
because homozygosity within couples, particularly at some HLA loci, is associated with modestly
reduced fertility [189,190]. Thus, leukocyte immunization was thought to introduce a needed
measure of immune recognition and stimulation. However, a multicenter study showed
conclusively that leukocyte immunization was not effective and strongly suggested that
reproductive outcomes might be worsened rather than improved by this intervention [191]. In
addition, a randomized clinical trial of intravenous immune globulin treatment in females with
repeated, unexplained in vitro fertilization failure found no improvement in the livebirth rate
[192]. A meta-analysis further confirmed that there is insufficient evidence to support the use of
these treatments for pregnancy losses [193]. (See "Recurrent pregnancy loss: Evaluation" and
"Recurrent pregnancy loss: Management".)
Other therapeutic possibilities for recurrent pregnancy loss include anticoagulants (low-
molecular-weight heparin, aspirin), hormones (progesterone), and immunomodulators
(granulocyte macrophage-colony stimulating factor [GM-CSF], human growth hormone [hCG],
macrophage-colony stimulating factor [M-CSF]) [194].

FETAL IMMUNE SYSTEM

Contributions of the fetal immune system to maternal-fetal immunologic interactions are not
well understood. Some evidence suggests that the fetal immune system is unable to mount an
anti-maternal immune response until mid- to late pregnancy. As an example, placental
macrophages do not express human leukocyte antigen (HLA) class II antigens until the second
trimester. Therefore, they are incapable of acting as fully effective antigen-presenting cells until
this time [195]. In contrast, another study showed that fetal T cells were highly responsive to
stimulation but were biased towards developing into regulatory T cells (Tregs), which are
important in tolerance [196]. However, if fetal T cells are able to mount a response to maternal
antigens, these fetal T cells may play a role in preterm labor [197].

Immunoglobulins — Immunoglobulin G (IgG) is transferred transplacentally from the mother


to the fetus during the third trimester, although other classes of immunoglobulins are not
under normal circumstances. Maternal IgG antibodies are usually beneficial to the infant, but
certain maternal antibodies may result in disease in the newborn infant, such as hemolytic
disease of the newborn (Rh incompatibility/Rh disease) [198]. Transplacental immunoglobulin
transfer is reviewed separately. (See "Placental development and physiology", section on
'Immunoglobulin G transfer'.)

FUTURE DIRECTIONS

Although much is now known of the immunologic conditions that lead to successful pregnancy,
much remains to be learned. Promising areas of research include the involvement of human
leukocyte antigen (HLA) G in angiogenesis and autoimmune disorders and the effect of
complement activation on dysregulation of angiogenesis leading to intrauterine fetal growth
restriction [199-201]. Utilization of new genomics, proteomics, and immunophenotyping
techniques may assist [112,202]. Immunomodulators and/or manipulation of innate immune
processes may also provide new avenues for therapeutics [194,203].

SUMMARY

● Prevention of immune rejection of the fetus requires local immunologic adaptations within
the mother, resulting in a state in which cytotoxic adaptive immune responses are
diminished, bypassed, or even abrogated while regulatory adaptive immunity is enhanced.
In contrast, innate (natural) immunity remains intact, serving two purposes: one, to
continue to provide host defense against infection, and two, to interact with fetal tissues to
promote successful placentation and pregnancy. (See 'Introduction' above.)

● Trophoblast cells protect the embryo itself and certain components of the extraembryonic
membranes. Multiple strategies are used by these cells to avoid maternal immune cells and
antibody-mediated cell destruction, including altered human leukocyte antigen (HLA)
expression, synthesis of immunoregulatory molecules, and expression of high levels of
complement regulatory proteins that protect the extraembryonic tissues from maternal
anti-paternal cytotoxic antibodies. (See 'Immune defense mechanisms of the placenta and
extraplacental membranes' above.)

● Uterine changes during pregnancy also help contribute to maternal immune adaptation,
including alterations in the relative proportions, phenotype, and functions of leukocyte
subpopulations, induction of immunoregulatory molecules (progesterone, prostaglandins),
and changes in cytokine profiles across gestation. (See 'Local immune adaptation in the
pregnant uterus' above.)

● Loss of potential pregnancies prior to or during implantation is common. Causes include


genetic and endocrine abnormalities and autoantibodies, such as antiphospholipid
antibodies (aPLs). Other potential causes are immune dysregulation and infection. Insults
such as infection and aPLs can alter normal placental function and mechanisms of immune
tolerance and can disrupt the normal trophoblast-maternal immune system crosstalk,
resulting in other adverse pregnancy outcomes, such as preterm labor and preeclampsia.
(See 'Immune factors in early pregnancy loss' above.)

ACKNOWLEDGMENT

The editorial staff at UpToDate would like to acknowledge E Richard Stiehm, MD, who
contributed as a Section Editor to an earlier version of this topic review.

Use of UpToDate is subject to the Subscription and License Agreement.

REFERENCES

1. Hunt JS. Stranger in a strange land. Immunol Rev 2006; 213:36.

2. Robertson SA. Immune regulation of conception and embryo implantation-all about quality
control? J Reprod Immunol 2010; 85:51.
3. Stoller M, Traupe T, Khattab AA, et al. Effects of coronary sinus occlusion on myocardial
ischaemia in humans: role of coronary collateral function. Heart 2013; 99:548.
4. Mor G, Abrahams VM. The immunology of pregnancy. In: Creasy and Resnik's Maternal-Fet
al Medicine: Principles and Practice, 7th ed., Creasy RK, Resnik R, Iams JD, et al (Eds), Elsevie
r, Philadelphia 2014. p.80.
5. Guerin LR, Prins JR, Robertson SA. Regulatory T-cells and immune tolerance in pregnancy: a
new target for infertility treatment? Hum Reprod Update 2009; 15:517.
6. Leber A, Teles A, Zenclussen AC. Regulatory T cells and their role in pregnancy. Am J Reprod
Immunol 2010; 63:445.
7. Barrientos G, Tirado-González I, Klapp BF, et al. The impact of dendritic cells on angiogenic
responses at the fetal-maternal interface. J Reprod Immunol 2009; 83:85.
8. Dekel N, Gnainsky Y, Granot I, Mor G. Inflammation and implantation. Am J Reprod
Immunol 2010; 63:17.
9. Nagamatsu T, Schust DJ. The contribution of macrophages to normal and pathological
pregnancies. Am J Reprod Immunol 2010; 63:460.
10. Zhang J, Chen Z, Smith GN, Croy BA. Natural killer cell-triggered vascular transformation:
maternal care before birth? Cell Mol Immunol 2011; 8:1.
11. Matzinger P. The danger model: a renewed sense of self. Science 2002; 296:301.

12. Mor G, Romero R, Aldo PB, Abrahams VM. Is the trophoblast an immune regulator? The
role of Toll-like receptors during pregnancy. Crit Rev Immunol 2005; 25:375.
13. Tong M, Abrahams VM. Immunology of the Placenta. Obstet Gynecol Clin North Am 2020;
47:49.
14. Gammill HS, Nelson JL. Naturally acquired microchimerism. Int J Dev Biol 2010; 54:531.
15. Frank HG. Placental development. In: Fetal and Neonatal Physiology, 5th ed, Polin RA, Abm
an SH, Rowitch D, Benitz WE (Eds), Elsevier, 2017. Vol 1, p.103.
16. Tilburgs T, Scherjon SA, Claas FH. Major histocompatibility complex (MHC)-mediated
immune regulation of decidual leukocytes at the fetal-maternal interface. J Reprod
Immunol 2010; 85:58.
17. Hackmon R, Pinnaduwage L, Zhang J, et al. Definitive class I human leukocyte antigen
expression in gestational placentation: HLA-F, HLA-E, HLA-C, and HLA-G in extravillous
trophoblast invasion on placentation, pregnancy, and parturition. Am J Reprod Immunol
2017; 77.
18. Ferreira LMR, Meissner TB, Tilburgs T, Strominger JL. HLA-G: At the Interface of Maternal-
Fetal Tolerance. Trends Immunol 2017; 38:272.
19. Petroff MG, Hunt JS. Immunity at the maternal-fetal interface. In: Mucosal immunology, 3rd
ed, Mestecky J, Lamm ME, Strober W, et al (Eds), Academic Press, New York 2005. p.1735.
20. Hunt JS, McIntire RH, Petroff MG. Immunobiology of human pregnancy. In: Knobil and Neil
l's physiology of reproduction, 3rd ed, Neill JD (Ed), Elsevier/Academic Press, St. Louis 2006.
Vol 2, p.2759.
21. Hunt JS, Langat DK, McIntire RH, Morales PJ. The role of HLA-G in human pregnancy.
Reprod Biol Endocrinol 2006; 4 Suppl 1:S10.

22. Long EO. Regulation of immune responses through inhibitory receptors. Annu Rev
Immunol 1999; 17:875.
23. Shakhawat A, Shaikly V, Elzatma E, et al. Interaction between HLA-G and
monocyte/macrophages in human pregnancy. J Reprod Immunol 2010; 85:40.
24. Gregori S, Amodio G, Quattrone F, Panina-Bordignon P. HLA-G Orchestrates the Early
Interaction of Human Trophoblasts with the Maternal Niche. Front Immunol 2015; 6:128.
25. Liu S, Diao L, Huang C, et al. The role of decidual immune cells on human pregnancy. J
Reprod Immunol 2017; 124:44.
26. Papúchová H, Meissner TB, Li Q, et al. The Dual Role of HLA-C in Tolerance and Immunity at
the Maternal-Fetal Interface. Front Immunol 2019; 10:2730.
27. Hiby SE, Walker JJ, O'shaughnessy KM, et al. Combinations of maternal KIR and fetal HLA-C
genes influence the risk of preeclampsia and reproductive success. J Exp Med 2004;
200:957.
28. Moffett A, Colucci F. Co-evolution of NK receptors and HLA ligands in humans is driven by
reproduction. Immunol Rev 2015; 267:283.
29. Tilburgs T, Meissner TB, Ferreira LMR, et al. NLRP2 is a suppressor of NF-ƙB signaling and
HLA-C expression in human trophoblasts†,‡. Biol Reprod 2017; 96:831.
30. Fournel S, Aguerre-Girr M, Huc X, et al. Cutting edge: soluble HLA-G1 triggers CD95/CD95
ligand-mediated apoptosis in activated CD8+ cells by interacting with CD8. J Immunol 2000;
164:6100.
31. Naji A, Durrbach A, Carosella ED, Rouas-Freiss N. Soluble HLA-G and HLA-G1 expressing
antigen-presenting cells inhibit T-cell alloproliferation through ILT-2/ILT-4/FasL-mediated
pathways. Hum Immunol 2007; 68:233.
32. Hunt JS, Jadhav L, Chu W, et al. Soluble HLA-G circulates in maternal blood during
pregnancy. Am J Obstet Gynecol 2000; 183:682.
33. Morales PJ, Pace JL, Platt JS, et al. Synthesis of beta(2)-microglobulin-free, disulphide-linked
HLA-G5 homodimers in human placental villous cytotrophoblast cells. Immunology 2007;
122:179.
34. Köstlin N, Ostermeir AL, Spring B, et al. HLA-G promotes myeloid-derived suppressor cell
accumulation and suppressive activity during human pregnancy through engagement of
the receptor ILT4. Eur J Immunol 2017; 47:374.
35. Lee CL, Guo Y, So KH, et al. Soluble human leukocyte antigen G5 polarizes differentiation of
macrophages toward a decidual macrophage-like phenotype. Hum Reprod 2015; 30:2263.
36. van der Meer A, Lukassen HG, van Cranenbroek B, et al. Soluble HLA-G promotes Th1-type
cytokine production by cytokine-activated uterine and peripheral natural killer cells. Mol
Hum Reprod 2007; 13:123.
37. Hunt JS, Fishback JL, Andrews GK, Wood GW. Expression of class I HLA genes by trophoblast
cells. Analysis by in situ hybridization. J Immunol 1988; 140:1293.
38. Hunt JS, Fishback JL, Chumbley G, Loke YW. Identification of class I MHC mRNA in human
first trimester trophoblast cells by in situ hybridization. J Immunol 1990; 144:4420.
39. Chu W, Fant ME, Geraghty DE, Hunt JS. Soluble HLA-G in human placentas: synthesis in
trophoblasts and interferon-gamma-activated macrophages but not placental fibroblasts.
Hum Immunol 1998; 59:435.
40. Murphy SP, Tomasi TB. Absence of MHC class II antigen expression in trophoblast cells
results from a lack of class II transactivator (CIITA) gene expression. Mol Reprod Dev 1998;
51:1.
41. Petroff MG, Chen L, Phillips TA, et al. B7 family molecules are favorably positioned at the
human maternal-fetal interface. Biol Reprod 2003; 68:1496.

42. Petroff MG, Perchellet A. B7 family molecules as regulators of the maternal immune
system in pregnancy. Am J Reprod Immunol 2010; 63:506.
43. Linscheid C, Petroff MG. Minor histocompatibility antigens and the maternal immune
response to the fetus during pregnancy. Am J Reprod Immunol 2013; 69:304.
44. Munn DH, Zhou M, Attwood JT, et al. Prevention of allogeneic fetal rejection by tryptophan
catabolism. Science 1998; 281:1191.
45. Ban Y, Zhao Y, Liu F, et al. Effect of Indoleamine 2,3-Dioxygenase Expressed in HTR-8/SVneo
Cells on Decidual NK Cell Cytotoxicity. Am J Reprod Immunol 2016; 75:519.
46. Hunt JS, Chen HL, Miller L. Tumor necrosis factors: pivotal components of pregnancy? Biol
Reprod 1996; 54:554.
47. Runic R, Lockwood CJ, Ma Y, et al. Expression of Fas ligand by human cytotrophoblasts:
implications in placentation and fetal survival. J Clin Endocrinol Metab 1996; 81:3119.
48. Phillips TA, Ni J, Pan G, et al. TRAIL (Apo-2L) and TRAIL receptors in human placentas:
implications for immune privilege. J Immunol 1999; 162:6053.

49. Phillips TA, Ni J, Hunt JS. Death-inducing tumour necrosis factor (TNF) superfamily ligands
and receptors are transcribed in human placentae, cytotrophoblasts, placental
macrophages and placental cell lines. Placenta 2001; 22:663.
50. Phillips TA, Ni J, Hunt JS. Cell-specific expression of B lymphocyte (APRIL, BLyS)- and Th2
(CD30L/CD153)-promoting tumor necrosis factor superfamily ligands in human placentas. J
Leukoc Biol 2003; 74:81.

51. Abrahams VM, Straszewski-Chavez SL, Guller S, Mor G. First trimester trophoblast cells
secrete Fas ligand which induces immune cell apoptosis. Mol Hum Reprod 2004; 10:55.
52. Frängsmyr L, Baranov V, Nagaeva O, et al. Cytoplasmic microvesicular form of Fas ligand in
human early placenta: switching the tissue immune privilege hypothesis from cellular to
vesicular level. Mol Hum Reprod 2005; 11:35.
53. Kshirsagar SK, Alam SM, Jasti S, et al. Immunomodulatory molecules are released from the
first trimester and term placenta via exosomes. Placenta 2012; 33:982.
54. Mincheva-Nilsson L, Baranov V. The role of placental exosomes in reproduction. Am J
Reprod Immunol 2010; 63:520.
55. Redman CW, Sargent IL. Circulating microparticles in normal pregnancy and pre-eclampsia.
Placenta 2008; 29 Suppl A:S73.
56. Holder BS, Tower CL, Forbes K, et al. Immune cell activation by trophoblast-derived
microvesicles is mediated by syncytin 1. Immunology 2012; 136:184.
57. Li H, Ouyang Y, Sadovsky E, et al. Unique microRNA Signals in Plasma Exosomes from
Pregnancies Complicated by Preeclampsia. Hypertension 2020; 75:762.
58. Delorme-Axford E, Donker RB, Mouillet JF, et al. Human placental trophoblasts confer viral
resistance to recipient cells. Proc Natl Acad Sci U S A 2013; 110:12048.
59. Holder B, Jones T, Sancho Shimizu V, et al. Macrophage Exosomes Induce Placental
Inflammatory Cytokines: A Novel Mode of Maternal-Placental Messaging. Traffic 2016;
17:168.
60. Denison FC, Kelly RW, Calder AA, Riley SC. Cytokine secretion by human fetal membranes,
decidua and placenta at term. Hum Reprod 1998; 13:3560.
61. Hunt JS. Cytokine networks in the human placenta. In: Cytokines in human reproduction, Hi
ll JA (Ed), Wiley-Liss, New York 1999. p.203.
62. Moreau P, Adrian-Cabestre F, Menier C, et al. IL-10 selectively induces HLA-G expression in
human trophoblasts and monocytes. Int Immunol 1999; 11:803.
63. Hennessy A, Pilmore HL, Simmons LA, Painter DM. A deficiency of placental IL-10 in
preeclampsia. J Immunol 1999; 163:3491.
64. Thaxton JE, Sharma S. Interleukin-10: a multi-faceted agent of pregnancy. Am J Reprod
Immunol 2010; 63:482.
65. Guo PF, Du MR, Wu HX, et al. Thymic stromal lymphopoietin from trophoblasts induces
dendritic cell-mediated regulatory TH2 bias in the decidua during early gestation in
humans. Blood 2010; 116:2061.
66. Hsi BL, Hunt JS, Atkinson JP. Differential expression of complement regulatory proteins on
subpopulations of human trophoblast cells. J Reprod Immunol 1991; 19:209.
67. Holmes CH, Simpson KL, Okada H, et al. Complement regulatory proteins at the feto-
maternal interface during human placental development: distribution of CD59 by
comparison with membrane cofactor protein (CD46) and decay accelerating factor (CD55).
Eur J Immunol 1992; 22:1579.
68. Xu C, Mao D, Holers VM, et al. A critical role for murine complement regulator crry in
fetomaternal tolerance. Science 2000; 287:498.

69. Koopman LA, Kopcow HD, Rybalov B, et al. Human decidual natural killer cells are a unique
NK cell subset with immunomodulatory potential. J Exp Med 2003; 198:1201.
70. Bulmer JN, Pace D, Ritson A. Immunoregulatory cells in human decidua: morphology,
immunohistochemistry and function. Reprod Nutr Dev 1988; 28:1599.
71. Gardner L, Moffett A. Dendritic cells in the human decidua. Biol Reprod 2003; 69:1438.
72. King A, Wellings V, Gardner L, Loke YW. Immunocytochemical characterization of the
unusual large granular lymphocytes in human endometrium throughout the menstrual
cycle. Hum Immunol 1989; 24:195.
73. Tilburgs T, Claas FH, Scherjon SA. Elsevier Trophoblast Research Award Lecture: Unique
properties of decidual T cells and their role in immune regulation during human
pregnancy. Placenta 2010; 31 Suppl:S82.
74. Ning F, Liu H, Lash GE. The Role of Decidual Macrophages During Normal and Pathological
Pregnancy. Am J Reprod Immunol 2016; 75:298.
75. Zhang YH, He M, Wang Y, Liao AH. Modulators of the Balance between M1 and M2
Macrophages during Pregnancy. Front Immunol 2017; 8:120.
76. Wang XQ, Zhou WJ, Hou XX, et al. Trophoblast-derived CXCL16 induces M2 macrophage
polarization that in turn inactivates NK cells at the maternal-fetal interface. Cell Mol
Immunol 2018; 15:1038.
77. Jiang X, Du MR, Li M, Wang H. Three macrophage subsets are identified in the uterus
during early human pregnancy. Cell Mol Immunol 2018; 15:1027.
78. Abrahams VM, Kim YM, Straszewski SL, et al. Macrophages and apoptotic cell clearance
during pregnancy. Am J Reprod Immunol 2004; 51:275.
79. Fest S, Aldo PB, Abrahams VM, et al. Trophoblast-macrophage interactions: a regulatory
network for the protection of pregnancy. Am J Reprod Immunol 2007; 57:55.
80. Harris LK. Review: Trophoblast-vascular cell interactions in early pregnancy: how to
remodel a vessel. Placenta 2010; 31 Suppl:S93.
81. Renaud SJ, Graham CH. The role of macrophages in utero-placental interactions during
normal and pathological pregnancy. Immunol Invest 2008; 37:535.
82. Eastabrook G, Brown M, Sargent I. The origins and end-organ consequence of pre-
eclampsia. Best Pract Res Clin Obstet Gynaecol 2011; 25:435.
83. Tang Z, Abrahams VM, Mor G, Guller S. Placental Hofbauer cells and complications of
pregnancy. Ann N Y Acad Sci 2011; 1221:103.
84. Tang Z, Buhimschi IA, Buhimschi CS, et al. Decreased levels of folate receptor-β and
reduced numbers of fetal macrophages (Hofbauer cells) in placentas from pregnancies
with severe pre-eclampsia. Am J Reprod Immunol 2013; 70:104.
85. Young OM, Tang Z, Niven-Fairchild T, et al. Toll-like receptor-mediated responses by
placental Hofbauer cells (HBCs): a potential pro-inflammatory role for fetal M2
macrophages. Am J Reprod Immunol 2015; 73:22.
86. Zhang J, Dunk C, Croy AB, Lye SJ. To serve and to protect: the role of decidual innate
immune cells on human pregnancy. Cell Tissue Res 2016; 363:249.
87. Kopcow HD, Allan DS, Chen X, et al. Human decidual NK cells form immature activating
synapses and are not cytotoxic. Proc Natl Acad Sci U S A 2005; 102:15563.
88. Croy BA, Luross JA, Guimond MJ, Hunt JS. Uterine natural killer cells: insights into lineage
relationships and functions from studies of pregnancies in mutant and transgenic mice.
Nat Immun 1997; 15:22.
89. Parr EL, Chen HL, Parr MB, Hunt JS. Synthesis and granular localization of tumor necrosis
factor-alpha in activated NK cells in the pregnant mouse uterus. J Reprod Immunol 1995;
28:31.
90. Stallmach T, Ehrenstein T, Isenmann S, et al. The role of perforin-expression by granular
metrial gland cells in pregnancy. Eur J Immunol 1995; 25:3342.
91. Rätsep MT, Felker AM, Kay VR, et al. Uterine natural killer cells: supervisors of vasculature
construction in early decidua basalis. Reproduction 2015; 149:R91.
92. Gamliel M, Goldman-Wohl D, Isaacson B, et al. Trained Memory of Human Uterine NK Cells
Enhances Their Function in Subsequent Pregnancies. Immunity 2018; 48:951.
93. Brighton PJ, Maruyama Y, Fishwick K, et al. Clearance of senescent decidual cells by uterine
natural killer cells in cycling human endometrium. Elife 2017; 6.
94. Hiby SE, Regan L, Lo W, et al. Association of maternal killer-cell immunoglobulin-like
receptors and parental HLA-C genotypes with recurrent miscarriage. Hum Reprod 2008;
23:972.
95. Management of high-risk pregnancy. In: Queenan's High-risk obstetrics, 6th ed, Queenan J
T, Spong CY, Lockwood CJ (Eds), Blackwell Publishing Ltd., Malden, MA 2011.
96. Moffett A, Hiby SE. How Does the maternal immune system contribute to the development
of pre-eclampsia? Placenta 2007; 28 Suppl A:S51.

97. Kwak-Kim J, Bao S, Lee SK, et al. Immunological modes of pregnancy loss: inflammation,
immune effectors, and stress. Am J Reprod Immunol 2014; 72:129.
98. Williams PJ, Bulmer JN, Searle RF, et al. Altered decidual leucocyte populations in the
placental bed in pre-eclampsia and foetal growth restriction: a comparison with late
normal pregnancy. Reproduction 2009; 138:177.
99. Fukui A, Funamizu A, Fukuhara R, Shibahara H. Expression of natural cytotoxicity receptors
and cytokine production on endometrial natural killer cells in women with recurrent
pregnancy loss or implantation failure, and the expression of natural cytotoxicity receptors
on peripheral blood natural killer cells in pregnant women with a history of recurrent
pregnancy loss. J Obstet Gynaecol Res 2017; 43:1678.
100. Lucas ES, Vrljicak P, Muter J, et al. Recurrent pregnancy loss is associated with a pro-
senescent decidual response during the peri-implantation window. Commun Biol 2020;
3:37.

101. Guimond MJ, Luross JA, Wang B, et al. Absence of natural killer cells during murine
pregnancy is associated with reproductive compromise in TgE26 mice. Biol Reprod 1997;
56:169.
102. Vacca P, Montaldo E, Croxatto D, et al. Identification of diverse innate lymphoid cells in
human decidua. Mucosal Immunol 2015; 8:254.
103. Croxatto D, Micheletti A, Montaldo E, et al. Group 3 innate lymphoid cells regulate
neutrophil migration and function in human decidua. Mucosal Immunol 2016; 9:1372.
104. Xu Y, Romero R, Miller D, et al. Innate lymphoid cells at the human maternal-fetal interface
in spontaneous preterm labor. Am J Reprod Immunol 2018; 79:e12820.
105. Krey G, Frank P, Shaikly V, et al. In vivo dendritic cell depletion reduces breeding efficiency,
affecting implantation and early placental development in mice. J Mol Med (Berl) 2008;
86:999.
106. Plaks V, Birnberg T, Berkutzki T, et al. Uterine DCs are crucial for decidua formation during
embryo implantation in mice. J Clin Invest 2008; 118:3954.
107. Laresgoiti-Servitje E, Gómez-López N, Olson DM. An immunological insight into the origins
of pre-eclampsia. Hum Reprod Update 2010; 16:510.
108. Collins MK, Tay CS, Erlebacher A. Dendritic cell entrapment within the pregnant uterus
inhibits immune surveillance of the maternal/fetal interface in mice. J Clin Invest 2009;
119:2062.
109. Tagliani E, Erlebacher A. Dendritic cell function at the maternal-fetal interface. Expert Rev
Clin Immunol 2011; 7:593.
110. Huang SJ, Chen CP, Schatz F, et al. Pre-eclampsia is associated with dendritic cell
recruitment into the uterine decidua. J Pathol 2008; 214:328.
111. Lockwood CJ, Matta P, Krikun G, et al. Regulation of monocyte chemoattractant protein-1
expression by tumor necrosis factor-alpha and interleukin-1beta in first trimester human
decidual cells: implications for preeclampsia. Am J Pathol 2006; 168:445.
112. Vazquez J, Chavarria M, Li Y, et al. Computational flow cytometry analysis reveals a unique
immune signature of the human maternal-fetal interface. Am J Reprod Immunol 2018; 79.
113. Heyborne KD, Cranfill RL, Carding SR, et al. Characterization of gamma delta T lymphocytes
at the maternal-fetal interface. J Immunol 1992; 149:2872.

114. Ito K, Karasawa M, Kawano T, et al. Involvement of decidual Valpha14 NKT cells in abortion.
Proc Natl Acad Sci U S A 2000; 97:740.

115. Mincheva-Nilsson L. Pregnancy and gamma/delta T cells: taking on the hard questions.
Reprod Biol Endocrinol 2003; 1:120.

116. Huang C, Zeng Y, Tu W. The role of γδ-T cells during human pregnancy. Am J Reprod
Immunol 2017; 78.

117. Tilburgs T, Strominger JL. CD8+ effector T cells at the fetal-maternal interface, balancing
fetal tolerance and antiviral immunity. Am J Reprod Immunol 2013; 69:395.
118. Zenclussen AC. Adaptive immune responses during pregnancy. Am J Reprod Immunol
2013; 69:291.
119. Lissauer D, Kilby MD, Moss P. Maternal effector T cells within decidua: The adaptive
immune response to pregnancy? Placenta 2017; 60:140.
120. Mjösberg J, Berg G, Jenmalm MC, Ernerudh J. FOXP3+ regulatory T cells and T helper 1, T
helper 2, and T helper 17 cells in human early pregnancy decidua. Biol Reprod 2010;
82:698.
121. Schumacher A, Zenclussen AC. Regulatory T cells: regulators of life. Am J Reprod Immunol
2014; 72:158.
122. Tilburgs T, Crespo ÂC, van der Zwan A, et al. Human HLA-G+ extravillous trophoblasts:
Immune-activating cells that interact with decidual leukocytes. Proc Natl Acad Sci U S A
2015; 112:7219.
123. Rowe JH, Ertelt JM, Xin L, Way SS. Pregnancy imprints regulatory memory that sustains
anergy to fetal antigen. Nature 2012; 490:102.
124. Kahn DA, Baltimore D. Pregnancy induces a fetal antigen-specific maternal T regulatory cell
response that contributes to tolerance. Proc Natl Acad Sci U S A 2010; 107:9299.
125. Sasaki Y, Darmochwal-Kolarz D, Suzuki D, et al. Proportion of peripheral blood and decidual
CD4(+) CD25(bright) regulatory T cells in pre-eclampsia. Clin Exp Immunol 2007; 149:139.
126. Williams Z. Inducing tolerance to pregnancy. N Engl J Med 2012; 367:1159.

127. Saito S, Nakashima A, Shima T, Ito M. Th1/Th2/Th17 and regulatory T-cell paradigm in
pregnancy. Am J Reprod Immunol 2010; 63:601.

128. Woidacki K, Zenclussen AC, Siebenhaar F. Mast cell-mediated and associated disorders in
pregnancy: a risky game with an uncertain outcome? Front Immunol 2014; 5:231.

129. Gomez-Lopez N, StLouis D, Lehr MA, et al. Immune cells in term and preterm labor. Cell
Mol Immunol 2014; 11:571.
130. Fettke F, Schumacher A, Costa SD, Zenclussen AC. B cells: the old new players in
reproductive immunology. Front Immunol 2014; 5:285.
131. Muzzio D, Zygmunt M, Jensen F. The role of pregnancy-associated hormones in the
development and function of regulatory B cells. Front Endocrinol (Lausanne) 2014; 5:39.
132. Lundell AC, Nordström I, Andersson K, et al. IFN type I and II induce BAFF secretion from
human decidual stromal cells. Sci Rep 2017; 7:39904.

133. Xu Y, Plazyo O, Romero R, et al. Isolation of Leukocytes from the Human Maternal-fetal
Interface. J Vis Exp 2015; :e52863.

134. Szekeres-Bartho J, Halasz M, Palkovics T. Progesterone in pregnancy; receptor-ligand


interaction and signaling pathways. J Reprod Immunol 2009; 83:60.

135. Szekeres-Bartho J, Wegmann TG. A progesterone-dependent immunomodulatory protein


alters the Th1/Th2 balance. J Reprod Immunol 1996; 31:81.

136. Miller L, Hunt JS. Regulation of TNF-alpha production in activated mouse macrophages by
progesterone. J Immunol 1998; 160:5098.
137. Wegmann TG, Lin H, Guilbert L, Mosmann TR. Bidirectional cytokine interactions in the
maternal-fetal relationship: is successful pregnancy a TH2 phenomenon? Immunol Today
1993; 14:353.

138. Lin H, Mosmann TR, Guilbert L, et al. Synthesis of T helper 2-type cytokines at the maternal-
fetal interface. J Immunol 1993; 151:4562.
139. Wei SQ, Fraser W, Luo ZC. Inflammatory cytokines and spontaneous preterm birth in
asymptomatic women: a systematic review. Obstet Gynecol 2010; 116:393.
140. Romero R, Mazor M, Brandt F, et al. Interleukin-1 alpha and interleukin-1 beta in preterm
and term human parturition. Am J Reprod Immunol 1992; 27:117.
141. Romero R, Mazor M, Sepulveda W, et al. Tumor necrosis factor in preterm and term labor.
Am J Obstet Gynecol 1992; 166:1576.

142. Mor G. Pregnancy reconceived: what keeps a mother's immune system from treating her b
aby as foreign tissue? A new theory resolves the paradox. Nat Hist 2007; 116:36. http://find
articles.com/p/articles/mi_m1134/is_4_116/ai_n19187208/ (Accessed on June 25, 2012).
143. Nancy P, Tagliani E, Tay CS, et al. Chemokine gene silencing in decidual stromal cells limits T
cell access to the maternal-fetal interface. Science 2012; 336:1317.

144. Salamonsen LA, Hannan NJ, Dimitriadis E. Cytokines and chemokines during human
embryo implantation: roles in implantation and early placentation. Semin Reprod Med
2007; 25:437.
145. Luppi P, Haluszczak C, Betters D, et al. Monocytes are progressively activated in the
circulation of pregnant women. J Leukoc Biol 2002; 72:874.
146. Amoudruz P, Minang JT, Sundström Y, et al. Pregnancy, but not the allergic status,
influences spontaneous and induced interleukin-1beta (IL-1beta), IL-6, IL-10 and IL-12
responses. Immunology 2006; 119:18.
147. Denney JM, Nelson EL, Wadhwa PD, et al. Longitudinal modulation of immune system
cytokine profile during pregnancy. Cytokine 2011; 53:170.
148. Veenstra van Nieuwenhoven AL, Bouman A, Moes H, et al. Endotoxin-induced cytokine
production of monocytes of third-trimester pregnant women compared with women in the
follicular phase of the menstrual cycle. Am J Obstet Gynecol 2003; 188:1073.
149. Sacks GP, Redman CW, Sargent IL. Monocytes are primed to produce the Th1 type cytokine
IL-12 in normal human pregnancy: an intracellular flow cytometric analysis of peripheral
blood mononuclear cells. Clin Exp Immunol 2003; 131:490.
150. Tafuri A, Alferink J, Möller P, et al. T cell awareness of paternal alloantigens during
pregnancy. Science 1995; 270:630.
151. Somerset DA, Zheng Y, Kilby MD, et al. Normal human pregnancy is associated with an
elevation in the immune suppressive CD25+ CD4+ regulatory T-cell subset. Immunology
2004; 112:38.
152. Aluvihare VR, Kallikourdis M, Betz AG. Regulatory T cells mediate maternal tolerance to the
fetus. Nat Immunol 2004; 5:266.
153. Zenclussen AC, Gerlof K, Zenclussen ML, et al. Abnormal T-cell reactivity against paternal
antigens in spontaneous abortion: adoptive transfer of pregnancy-induced CD4+CD25+ T
regulatory cells prevents fetal rejection in a murine abortion model. Am J Pathol 2005;
166:811.

154. Schumacher A, Wafula PO, Bertoja AZ, et al. Mechanisms of action of regulatory T cells
specific for paternal antigens during pregnancy. Obstet Gynecol 2007; 110:1137.

155. Chaouat G, Assal Meliani A, Martal J, et al. IL-10 prevents naturally occurring fetal loss in
the CBA x DBA/2 mating combination, and local defect in IL-10 production in this abortion-
prone combination is corrected by in vivo injection of IFN-tau. J Immunol 1995; 154:4261.

156. Svensson L, Arvola M, Sällström MA, et al. The Th2 cytokines IL-4 and IL-10 are not crucial
for the completion of allogeneic pregnancy in mice. J Reprod Immunol 2001; 51:3.

157. Rolle L, Memarzadeh Tehran M, Morell-García A, et al. Cutting edge: IL-10-producing


regulatory B cells in early human pregnancy. Am J Reprod Immunol 2013; 70:448.

158. Bardos J, Hercz D, Friedenthal J, et al. A national survey on public perceptions of


miscarriage. Obstet Gynecol 2015; 125:1313.
159. Norwitz ER. Defective implantation and placentation: laying the blueprint for pregnancy
complications. Reprod Biomed Online 2006; 13:591.
160. Robertson SA, O'Leary S, Armstrong DT. Influence of semen on inflammatory modulators
of embryo implantation. Soc Reprod Fertil Suppl 2006; 62:231.
161. Moffett A, Regan L, Braude P. Natural killer cells, miscarriage, and infertility. BMJ 2004;
329:1283.

162. De Carolis C, Perricone C, Perricone R. NK cells, autoantibodies, and immunologic infertility:


a complex interplay. Clin Rev Allergy Immunol 2010; 39:166.
163. Huang C, Zhang H, Chen X, et al. Association of peripheral blood dendritic cells with
recurrent pregnancy loss: a case-controlled study. Am J Reprod Immunol 2016; 76:326.

164. Beaman KD, Ntrivalas E, Mallers TM, et al. Immune etiology of recurrent pregnancy loss
and its diagnosis. Am J Reprod Immunol 2012; 67:319.

165. Dahl M, Hviid TV. Human leucocyte antigen class Ib molecules in pregnancy success and
early pregnancy loss. Hum Reprod Update 2012; 18:92.

166. Triebwasser MP, Wu X, Bertram P, et al. Timing and mechanism of conceptus demise in a
complement regulatory membrane protein deficient mouse. Am J Reprod Immunol 2018;
80:e12997.

167. Regal JF, Gilbert JS, Burwick RM. The complement system and adverse pregnancy
outcomes. Mol Immunol 2015; 67:56.

168. Jin LP, Fan DX, Zhang T, et al. The costimulatory signal upregulation is associated with Th1
bias at the maternal-fetal interface in human miscarriage. Am J Reprod Immunol 2011;
66:270.

169. Banzato PC, Daher S, Traina E, et al. FAS and FAS-L genotype and expression in patients
with recurrent pregnancy loss. Reprod Sci 2013; 20:1111.

170. Martius J, Eschenbach DA. The role of bacterial vaginosis as a cause of amniotic fluid
infection, chorioamnionitis and prematurity--a review. Arch Gynecol Obstet 1990; 247:1.

171. Gibbs RS, Romero R, Hillier SL, et al. A review of premature birth and subclinical infection.
Am J Obstet Gynecol 1992; 166:1515.
172. Cardenas I, Means RE, Aldo P, et al. Viral infection of the placenta leads to fetal
inflammation and sensitization to bacterial products predisposing to preterm labor. J
Immunol 2010; 185:1248.

173. Abrahams VM. Pattern recognition at the maternal-fetal interface. Immunol Invest 2008;
37:427.
174. Abrahams VM. The role of the Nod-like receptor family in trophoblast innate immune
responses. J Reprod Immunol 2011; 88:112.
175. Cervera R, Balasch J. Autoimmunity and recurrent pregnancy losses. Clin Rev Allergy
Immunol 2010; 39:148.
176. Girardi G, Berman J, Redecha P, et al. Complement C5a receptors and neutrophils mediate
fetal injury in the antiphospholipid syndrome. J Clin Invest 2003; 112:1644.

177. Girardi G, Redecha P, Salmon JE. Heparin prevents antiphospholipid antibody-induced fetal
loss by inhibiting complement activation. Nat Med 2004; 10:1222.

178. Berman J, Girardi G, Salmon JE. TNF-alpha is a critical effector and a target for therapy in
antiphospholipid antibody-induced pregnancy loss. J Immunol 2005; 174:485.

179. Redecha P, Tilley R, Tencati M, et al. Tissue factor: a link between C5a and neutrophil
activation in antiphospholipid antibody induced fetal injury. Blood 2007; 110:2423.

180. Tincani A, Cavazzana I, Ziglioli T, et al. Complement activation and pregnancy failure. Clin
Rev Allergy Immunol 2010; 39:153.
181. Girardi G. Role of tissue factor in the maternal immunological attack of the embryo in the
antiphospholipid syndrome. Clin Rev Allergy Immunol 2010; 39:160.
182. Cohen H, Cuadrado MJ, Erkan D, et al. 16th International Congress on Antiphospholipid
Antibodies Task Force Report on Antiphospholipid Syndrome Treatment Trends. Lupus
2020; 29:1571.
183. Gustavsen A, Skattum L, Bergseth G, et al. Effect on mother and child of eculizumab given
before caesarean section in a patient with severe antiphospholipid syndrome: A case
report. Medicine (Baltimore) 2017; 96:e6338.
184. Mulla MJ, Brosens JJ, Chamley LW, et al. Antiphospholipid antibodies induce a pro-
inflammatory response in first trimester trophoblast via the TLR4/MyD88 pathway. Am J
Reprod Immunol 2009; 62:96.

185. Mulla MJ, Myrtolli K, Brosens JJ, et al. Antiphospholipid antibodies limit trophoblast
migration by reducing IL-6 production and STAT3 activity. Am J Reprod Immunol 2010;
63:339.

186. Carroll TY, Mulla MJ, Han CS, et al. Modulation of trophoblast angiogenic factor secretion by
antiphospholipid antibodies is not reversed by heparin. Am J Reprod Immunol 2011;
66:286.

187. Mulla MJ, Salmon JE, Chamley LW, et al. A role for uric acid and the Nalp3 inflammasome in
antiphospholipid antibody-induced IL-1β production by human first trimester trophoblast.
PLoS One 2013; 8:e65237.
188. Mowbray JF, Gibbings C, Liddell H, et al. Controlled trial of treatment of recurrent
spontaneous abortion by immunisation with paternal cells. Lancet 1985; 1:941.

189. Ober C, Hyslop T, Hauck WW. Inbreeding effects on fertility in humans: evidence for
reproductive compensation. Am J Hum Genet 1999; 64:225.

190. Ober C, Hyslop T, Elias S, et al. Human leukocyte antigen matching and fetal loss: results of
a 10 year prospective study. Hum Reprod 1998; 13:33.

191. Ober C, Karrison T, Odem RR, et al. Mononuclear-cell immunisation in prevention of


recurrent miscarriages: a randomised trial. Lancet 1999; 354:365.

192. Stephenson MD, Fluker MR. Treatment of repeated unexplained in vitro fertilization failure
with intravenous immunoglobulin: a randomized, placebo-controlled Canadian trial. Fertil
Steril 2000; 74:1108.

193. Rasmark Roepke E, Hellgren M, Hjertberg R, et al. Treatment efficacy for idiopathic
recurrent pregnancy loss - a systematic review and meta-analyses. Acta Obstet Gynecol
Scand 2018; 97:921.

194. Matthiesen L, Kalkunte S, Sharma S. Multiple pregnancy failures: an immunological


paradigm. Am J Reprod Immunol 2012; 67:334.

195. Hunt JS, Robertson SA. Uterine macrophages and environmental programming for
pregnancy success. J Reprod Immunol 1996; 32:1.

196. Mold JE, Michaëlsson J, Burt TD, et al. Maternal alloantigens promote the development of
tolerogenic fetal regulatory T cells in utero. Science 2008; 322:1562.
197. Frascoli M, Coniglio L, Witt R, et al. Alloreactive fetal T cells promote uterine contractility in
preterm labor via IFN-γ and TNF-α. Sci Transl Med 2018; 10.
198. Hoftman AC, Hernandez MI, Lee KW, Stiehm ER. Newborn illnesses caused by
transplacental antibodies. Adv Pediatr 2008; 55:271.
199. Fons P, Chabot S, Cartwright JE, et al. Soluble HLA-G1 inhibits angiogenesis through an
apoptotic pathway and by direct binding to CD160 receptor expressed by endothelial cells.
Blood 2006; 108:2608.
200. Nicolae D, Cox NJ, Lester LA, et al. Fine mapping and positional candidate studies identify
HLA-G as an asthma susceptibility gene on chromosome 6p21. Am J Hum Genet 2005;
76:349.
201. Girardi G, Yarilin D, Thurman JM, et al. Complement activation induces dysregulation of
angiogenic factors and causes fetal rejection and growth restriction. J Exp Med 2006;
203:2165.

202. Buhimschi CS, Weiner CP, Buhimschi IA. Proteomics, part II: the emerging role of
proteomics over genomics in spontaneous preterm labor/birth. Obstet Gynecol Surv 2006;
61:543.

203. Abrahams VM. Antagonizing toll-like receptors to prevent preterm labor. Reprod Sci 2008;
15:108.
Topic 3966 Version 23.0
GRAPHICS

Embryo and the trophoblast at the end of the third week of development

Presomite embryo and the trophoblast at the end of the third week. Tertiary and secondary stem villi give the trophoblast
a characteristic radial appearance. Intervillous spaces, which are found throughout the trophoblast, are lined with
syncytium. Cytotrophoblastic cells surround the trophoblast entirely and are in direct contact with the endometrium. The
embryo is suspended in the chorionic cavity by means of the connecting stalk.

Reproduced with permission from: Sadler TW. Langman's Medical Embryology, 13th Edition, Lippincott Williams & Wilkins, Philadelphia
2015. Copyright © 2015 Lippincott Williams & Wilkins. www.lww.com.

Graphic 96332 Version 3.0


Contributor Disclosures
Vikki M Abrahams, PhD Nothing to disclose Rebecca Marsh, MD Nothing to disclose Charles J
Lockwood, MD, MHCM Nothing to disclose Elizabeth TePas, MD, MS Nothing to disclose

Contributor disclosures are reviewed for conflicts of interest by the editorial group. When found, these are
addressed by vetting through a multi-level review process, and through requirements for references to be
provided to support the content. Appropriately referenced content is required of all authors and must
conform to UpToDate standards of evidence.

Conflict of interest policy

You might also like