You are on page 1of 8

BBA - Molecular Basis of Disease 1866 (2020) 165349

Contents lists available at ScienceDirect

BBA - Molecular Basis of Disease


journal homepage: www.elsevier.com/locate/bbadis

Biology of preeclampsia: Combined actions of angiogenic factors, their T


receptors and placental proteins☆
Berthold Huppertz

Division of Cell Biology, Histology and Embryology, Gottfried Schatz Research Center, Medical University of Graz, Graz, Austria

ARTICLE INFO ABSTRACT

Keywords: Although massive efforts have been undertaken to elucidate the etiology of the pregnancy syndrome pre-
Preeclampsia eclampsia, its developmental origin remains a mystery. Most efforts of the last decade have focused on bio-
PP13 markers to predict and/or diagnose preeclampsia, including the anti-angiogenic factor sFlt-1 (soluble fms-like
PGF tyrosin kinase-1), the angiogenic factor PGF (placental growth factor) and PP13 (placental protein 13). The
sFlt-1
origins of these marker proteins are still under debate, and so far their actions have only been describe separate
Sepsis
Vasodilation
from each other. This study will focus on the origins and actions of all three markers during pregnancy and
outside pregnancy and will describe a scenario where all three markers act synergistically to rescue the mother
from the deleterious effects of the debris that is released from the placenta during preeclampsia. This more
holistic approach may open new avenues to think about maternal-fetal interactions and putative therapies.

1. Introduction monitoring the progress of the syndrome. The task force for hyperten-
sion in pregnancy of the American College of Obstetricians and
Preeclampsia, the syndrome of hypotheses, remains one of the Gynecologists (ACOG) clearly defined in 2013 [2]: “Screening to predict
major pathologies of pregnancy that results in mortality and morbidity preeclampsia beyond obtaining an appropriate medical history to evaluate
of mothers, fetuses and neonates. Recent data from the USA show that for risk factors is not recommended.”. Only regarding options of ther-
the rate of preeclampsia in inpatient deliveries increased by 21% from apeutic approaches, a first break through has been made. Based on data
2005 to 2014 [1]. In 2014 a total of 4.7% of all inpatient deliveries in of the ASPRE study, low dose aspirin given daily from 11–14 weeks of
the USA were associated with preeclampsia, i.e. 176,925 women. Also, pregnancy to 36 weeks lowers the rate of preterm preeclampsia sig-
looking at the numbers from 2014, 37% of all women suffering from nificantly [3].
preeclampsia showed clinical features of the severe form of the syn- This overview will focus on specific biomarkers that have been
drome, which is an increase of 50% from 2005 to 2014. At the same targets of extensive studies and will discuss their involvement in recent
time, the percentage of cases developing eclampsia slightly dropped developments in explaining the etiology of preeclampsia. Interestingly,
from 0.9% to 0.7%. Interestingly, the numbers from the USA show that as described above, the quest to identify predictive biomarkers for
black women have a rate of preeclampsia of 7.0%, while only 4.3% of preeclampsia already lasts two decades. At the same time, the WHO has
white women developed the syndrome [1]. The above numbers clearly defined requirements for biomarkers in general [4], while none of the
show the impact of preeclampsia on pregnant women and their children predictive biomarkers for preeclampsia identified so far meets these
even today and even in industrialized countries. requirements. According to the WHO, the biomarker (1) should be
Based on this impact, the last decades have seen massive efforts to specific for a well-defined disease with a known prevalence, (2) should
elucidate the etiology of preeclampsia, to identify predictive bio- be used with a disease where the respective information of the bio-
markers and to start preventive therapeutic approaches. Today, new marker can improve outcome and decrease severity, and (3) should be
approaches to identify putative predicting biomarkers include disease cost effective.
map analysis and circulating RNA markers.
So far, none of these efforts has resulted in unravelling the etiology
of preeclampsia nor in the clinical use of a single biomarker for


This article is part of a Special Issue entitled: Membrane Transporters and Receptors in Pregnancy Metabolic Complications edited by Luis Sobrevia.

Division of Cell Biology, Histology and Embryology, Gottfried Schatz Research Center, Medical University of Graz, Neue Stiftingtalstrasse 6/II, 8010 Graz, Austria.
E-mail address: berthold.huppertz@medunigraz.at.

https://doi.org/10.1016/j.bbadis.2018.11.024
Received 21 September 2018; Received in revised form 7 November 2018; Accepted 22 November 2018
Available online 13 December 2018
0925-4439/ © 2018 Elsevier B.V. All rights reserved.
B. Huppertz BBA - Molecular Basis of Disease 1866 (2020) 165349

2. Definition of preeclampsia 2.3. Evolving hypotheses on the etiology of preeclampsia

2.1. Diagnosis Preeclampsia is a pregnancy specific syndrome. The current hy-


potheses describe this syndrome to be centered on the presence of a
The diagnosis of preeclampsia is based on changes in blood pressure placenta. The release of so far unknown factors from the placenta over
and kidney function in a so far normotensive pregnant woman after the course of pregnancy finally leads to the occurrence of the clinical
20 weeks of pregnancy [5]. Values of ≥140 mm Hg for systolic blood symptoms of preeclampsia of the mother [9]. Of course, the suscept-
pressure and/or ≥90 mm Hg for diastolic blood pressure indicate the ibility of the mother is a major player in the development of this syn-
development of hypertension, while values of ≥300 mg/day protein in drome [9,10]. Hence, it is believed to be the interaction between the
the urine or a protein to creatinine ratio of ≥3 indicate the develop- quality and quantity of placental factors and how the mother can
ment of proteinuria. Recently, the fact that preeclampsia is a syndrome handle these factors that decides whether a woman develops pre-
rather than a defined disease led to the widening of the diagnosis and eclampsia or remains healthy. The combination of placental factors and
the inclusion of other criteria that indicate dysfunctions of other organs the maternal susceptibility seems to define the time of onset and the
such as liver, kidney, lungs and brain [5]. severity of symptoms as well as the progression of the syndrome and the
association with fetal growth [9,10]. Of course, it is still a matter of
discussion whether early and late onset types of preeclampsia share the
2.2. Subtypes same etiology or derive from mechanistically different defects [9].
Very recently, a new hypothesis has been developed centering on
80% to 90% of all preeclampsia cases are pure maternal syndromes the pregnant woman alone [11]. This new hypothesis is based on the
with normally grown babies. As an example, of the 26,893 deliveries in cardiovascular predisposition of the woman to preeclampsia. If the
a single hospital within a two year period, 7.3% were diagnosed with woman's cardiovascular system fails to adapt to the changes occurring
preeclampsia, of which 92.1% delivered after 34 weeks of pregnancy. In in pregnancy, this may well lead to the symptoms of preeclampsia. As
the group with preeclampsia 24.4% of all cases were delivered with a has been described before [9,10], preeclampsia may develop with a
small for gestational age fetus, while in the non-preeclampsia group very normal placenta but a highly susceptible mother. The recent hy-
10.8% of all fetuses were born small for gestational age [6]. In the pothesis [11] focuses on the susceptibility of the maternal cardiovas-
majority of these cases delivery takes place at term or slightly preterm, cular system and its adaptation to pregnancy. In this new concept,
while a small percentage of cases is additionally associated with growth factors released from the placenta are not decisive for the development
restriction of the baby (intra uterine growth restriction, IUGR). Most of of preeclampsia, but rather the ability of the maternal cardiovascular
these combined cases belong to the type of early-onset preeclampsia system to adapt to the changes occurring during pregnancy. Further
with the need for delivery prior to 34 weeks of gestation. In addition, knowledge is needed to unravel whether placental factors are needed
such combined cases affect both, mother and child, and hence are those for the development of preeclampsia or whether it is a pure mala-
of the highest clinical relevance. daptation of the woman to vascular changes during pregnancy [10,11].
The above differences have led to a sub-classification of pre-
eclampsia into different subtypes [7]: 3. Placental growth factor (PGF) and its soluble receptor sFlt-1

- Early-onset preeclampsia with delivery prior to 34 weeks of gesta- 3.1. Expression pattern in the placenta
tion
- Late-onset preeclampsia with delivery after 34 weeks of gestation Placental growth factor (PGF) and its soluble receptor (soluble fms-
like tyrosin kinase-1, sFlt-1) are well known predictive markers for
In another recent study [8], 35,948 singleton pregnancies were in- adverse pregnancy outcome including but not limited to preeclampsia
cluded, of which 1058 (2.9%) developed preeclampsia. These authors [12]. Both have been detected to be expressed in the placenta 20 years
used a different cut-off and defined term preeclampsia (delivery at > ago [13,14]. Expression in the placenta was mainly found in the villous
37 weeks of gestation) compared to preterm preeclampsia (delivery syncytiotrophoblast, the cover of the placental villi in direct contact
at < 37 weeks of gestation). The weeks of delivery for the preeclampsia with maternal blood. Release of PGF and sFlt-1 from this tissue directly
cases were as follows: 6.2% at < 32 weeks, 21.4% at 32 + 0 to flushes the angiogenic factor PGF and its soluble scavenging receptor
36 + 6 weeks, 48.6% at 37 + 0 to 39 + 6 weeks and 23.8% at > 40 sFlt-1 into the maternal circulation with actions on cells of the maternal
weeks. Hence, 72.4% of all preeclampsia cases delivered after 37 weeks vascular system.
of gestation [8]. Early reports showed that placental expression of PGF mRNA was
Interestingly, since pregnancy is terminated at 40 + 6 weeks in most slightly decreased at the end of the first trimester [15] and unchanged
industrialized countries today, those cases that would have developed at the time of symptoms [16], while sFlt-1/Flt-1 mRNA expression was
term preeclampsia at 41 or 42 weeks of gestation are no longer included upregulated at both time points [15,16]. In maternal blood, total (free
in the percentage values above. Hence, the proportion of late-onset and bound) VEGF, the natural growth factor for the membrane bound
preeclampsia cases may even be higher than reported here. Also, taking Flt-1 receptor, was increased, while free PGF was decreased and sFlt-1,
into account the above regimen of termination of pregnancy at the soluble component of the VEGF receptor family, was increased in
40 + 6 weeks, the performance of predictive markers such as mean cases with preeclampsia [16].
arterial pressure (MAP), which showed the best performance as a single In contrast to such findings, the premature conclusion was drawn
predictive marker in this study [8], may be underestimated as a certain that (1) the placenta is the sole source of these factors [17,18], and (2)
number of cases of preeclampsia (> 41 weeks) are no longer reaching the placenta-derived factors are nearly exclusively decisive for their
disease state. concentrations in maternal blood [19–21].
There is a variety of other definitions such as preterm, intermediate Recently, this topic has been revisited combining maternal serum
and early and late preterm, all subclassifying preeclampsia based on and placental tissues from the same pregnancies [22,23]. These authors
gestational age at delivery. There are also definitions based on the se- showed that reduced maternal PGF serum concentrations in pre-
verity of symptoms leading to subtypes such as mild and severe pre- eclampsia cases were associated with unchanged levels of PGF mRNA
eclampsia. and protein in the placenta as a whole. Interestingly, looking more
specifically, the PGF protein expression in the syncytiotrophoblast was
significantly increased in late onset preeclampsia cases [22], while it

2
B. Huppertz BBA - Molecular Basis of Disease 1866 (2020) 165349

remained unchanged in cases with early onset preeclampsia [23] Second trimester (19 to 24 weeks): IUGR is best predicted using a
compared to age-matched controls. These data shows that a simple link marker panel including serum PGF, serum alpha-fetoprotein, fetal
between changes of placental expression and respective changes in biometry and maternal risk factors with [42] or without uterine artery
maternal serum cannot easily be made. Thus, the serum levels of this pulsatility index [43]. The AUC value to predict early-onset IUGR was
factor and its soluble receptor need to be explained taking into account 0.999 in both publications. The comparison of PGF with other serum
maternal sources [24,25]. markers revealed that only PGF showed significantly different levels
If other sources are important in adding to the serum concentrations between controls and IUGR cases, while other factors including sFlt-1
of PGF and sFlt-1, the surface of the placental villous tree in comparison did not show any differences between cases and controls [43].
to the surface of the maternal vascular system becomes a decisive ele- Third trimester: The sFlt-1/PGF ratio has been shown to be very
ment. The surface of the villous tree is much smaller than the surface of helpful in predicting adverse outcome of pregnancy, hence not speci-
the vascular system of the mother. The maximal surface of the placental fically focusing on preeclampsia [44]. Looking at early-onset pre-
villous tree at term is about 12 to 15 m2 [26], while the surface of an eclampsia, the sFlt-1/PGF ratio has a clear advantage in patients who
adult vascular system has an area of > 3000 m2 [27]. Considering the have already been diagnosed with preeclampsia in predicting time to
release of a factor with a similar range from the placenta and the ma- delivery within a two or even four week period [12,44].
ternal endothelium, the placenta-derived factor would contribute to the
total serum concentration with only 0.4% to 0.5%. Thus, even major 3.4. Use of PGF and sFlt-1 outside pregnancy
changes of the placental release will change the total serum con-
centration only marginally. These data shows that maternal sources of PGF and Flt-1 are angiogenic factors that are systemically expressed
factors and receptors need to be considered [25]. in the vascular system to control angiogenesis [36]. With this expres-
Importantly, the false presumption that angiogenic and anti-angio- sion pattern, they have been good candidates for the prediction of
genic factors are altered in maternal serum of preeclamptic women due diseases. In fact, PGF is used as a marker for colorectal and ovarian
to placental hypoxia has already been disproved [28]. These authors cancer [45–47], while the PGF/sFlt-1 ratio is used for the stratification
compared levels of PGF and sFlt-1 in preeclamptic and healthy pregnant of e.g. breast cancer patients [48].
women at altitudes of 400 m and 3600 m and did not find any differ- Also, sFlt-1 alone has been used as a marker for subarachnoid he-
ences between low and high altitude (i.e. normoxia and hypoxia, re- morrhage hypoxia [49], and studies have been published evaluating the
spectively). This study reveals that oxygenation of maternal blood and use of sFlt-1 for cancer treatment [50]. This anti-angiogenic soluble
hence the placenta does not have any effect on the levels of PGF and receptor is known to be increased in sepsis and has been described as a
sFlt-1 in maternal blood [28]. marker to identify severity stages of sepsis [36,51]. Similar data has
been published for cases with acute pancreatitis where sFlt-1 is used to
3.2. Expression pattern outside the placenta predict the severity of symptoms [52].
A direct comparison of changes and serum concentrations of PGF
Of course, the placenta is not the only source of PGF and sFlt-1 and sFlt-1 in preeclampsia [53] and plasma cell myeloma [54] revealed
[24,25]. While the expression of PGF in quiescent endothelial cells is that in both cases not only the changes are going into the same direction
low, activated and/or stressed endothelial cells upregulate the expres- but also the serum concentrations are very similar [25]. This compar-
sion and secretion of PGF [29]. Similarly, Flt-1 is expressed in en- ison is important as it shows that the serum levels of PGF and sFlt-1 as
dothelial cells throughout the whole vascular system. Both, PGF and found in preeclampsia are also measured in diseases of non-pregnant
Flt-1 are further expressed in non-endothelial cells such as fibroblasts humans.
[30,31], osteoblasts [32], monocytes [33] and smooth muscle cells Applying this knowledge to preeclampsia and the placenta, one
[34]. The soluble Flt-1 receptor has been shown to be involved in could argue that non-angiogenic placental factors activate the maternal
vascular changes of the cycling endometrium [35]. A recent review endothelium, which in turn alters its secretion pattern of angiogenic
highlights the importance of VEGF/PGF and their receptors and de- factors such as PGF and sFlt-1.
scribes their body-wide expression pattern [36]. These authors describe Different to what has been speculated at the time the angiogenic
the presence of sFlt-1, but also of other soluble forms of the VEGF re- factors were detected as markers for preeclampsia, it has become clear
ceptor-1, namely sFlt1-14 (or sFlt1-e5a) and an additional soluble form today that:
of Flt1 generated by cleavage of the transmembrane part of the protein.
All soluble Fllt-1 receptors act as antagonists to the transmembrane Flt- (1) (Anti-) angiogenic factors are not specifically expressed in the pla-
1 receptor. The authors also point to the fact that sFlt-1 levels in blood centa but rather systemically in the mother/adult.
are directly linked to cell function of endothelial cells in the kidney (2) These factors are not specific markers to predict preeclampsia but
leading to increased blood levels of sFlt-1 in patients with chronic rather can be used to predict adverse outcome in a variety of syn-
kidney disease [36]. dromes and diseases during pregnancy and outside pregnancy.
(3) These factors are valuable markers to predict time to delivery once
3.3. Use of PGF and sFlt-1 in pregnancy an adverse outcome during pregnancy has been diagnosed.

First trimester (11 to 13 weeks): Early in pregnancy, the serum level 3.5. Actions of PGF and sFlt-1 in general
of sFlt-1 does not show any significant changes in preeclampsia com-
pared to healthy pregnancies [37–39]. In addition, a very recent direct As an angiogenic factor, PGF is not only involved in growth and
comparison of first trimester angiogenic factor levels revealed no maturation of new vessels from already existing vessels [55], but also
changes of VEGF, PGF, sFlt-1 and sEng in late onset preeclampsia versus shows direct action on existing vessels [56]. PGF directly interacts with
controls [40]. endothelial cells and enhances their survival/viability as well as pro-
However, early-onset preeclampsia as well as IUGR can be equally liferation and migration [55,57]. Moreover, PGF induces dilation of
well predicted using a set of first trimester biomarkers including serum resistance arteries in the uterus, the mesentery and the skin [58].
PGF, the uterine artery pulsatility index and maternal risk factors with Interestingly, PGF seems to be more active in disease states rather
areas under the curve (AUC) of 0.92 for early-onset preeclampsia and than in health and development: PGF deficient mice are healthy, viable,
0.91 for IUGR [41]. The prediction of late-onset preeclampsia is best and fertile and display a normal development [57]. At the same time,
done without serum PGF, but rather taking into account mean arterial PGF has been reported to be up or downregulated in a large variety of
blood pressure and maternal risk factors [41]. diseases outside pregnancy as well as preeclampsia [56]. It has been

3
B. Huppertz BBA - Molecular Basis of Disease 1866 (2020) 165349

hypothesized that alterations of PGF in disease states may be used to twice or three times as high as the level in the first trimester [69].
counteract disease-related changes while this does not interfere with Interestingly, PP13 can still be detected in maternal serum two to five
normal physiology [56]. weeks after delivery, in the absence of the placenta as the only source
In addition, sFlt-1 has been described to be increased in cases of [69].
sepsis [36,51]. Shapiro et al. [51] speculated on the function of sFlt-1 to In preeclampsia, the maternal serum levels show a biphasic beha-
sequester the high levels of VEGF is these cases, thereby reducing the vior. During the first and early second trimester, the levels are sig-
negative effects of VEGF in sepsis such as vessel leakage. VEGF has been nificantly lower compared to healthy controls [69,73]. At the time of
shown to promote internalization of VE-cadherin leading to dis- clinical symptoms of preeclampsia, the PP13 levels in maternal serum
assembly of intercellular junctions of endothelial cells and thus vessel rise above normal levels up to twice the level compared to controls
leakage [59]. In a computational model, the interaction between VEGF [69]. In vitro, villous explants from placentas of preeclamptic women
and sFlt-1 has been calculated [60]. These authors could show that release much higher amounts of PP13 compared to age-matched con-
augmented local concentrations of sFlt-1 lead to local sequestration of trols [71].
VEGF and reduced signaling through the transmembrane receptors Flt-1 Today, the use of PP13 as a predictive biomarker for preeclampsia
and Flk-1 [60]. derived from maternal serum has been investigated in over 20 studies
(e.g. [69,72–76]). Unfortunately, various assays and platforms have
4. Placental protein 13 been used over time, which had a clearly negative effect on the per-
formance of the marker [77]. Nevertheless, testing of PP13 in different
4.1. Expression pattern in the placenta cohorts and various risk populations revealed a detection rate for PP13
as single marker of 83% for early-onset preeclampsia (delivery prior to
Placental protein 13 (PP13) RNA is expressed in the amnion of the 34 weeks), 66% for preterm preeclampsia (delivery prior to 37 weeks)
fetal membranes as well as in the placental syncytiotrophoblast [61]. and 47% for all preeclampsia cases at a false positive rate of 10% [77].
The expression of PP13 is nearly exclusive to the fetal tissues of the
placenta, while other fetal and adult tissues show very little to no ex- 4.3. Actions of PP13 during and outside pregnancy
pression of this protein [61]. Recently, it has been shown that micro-
vesicles and exosomes, released from the syncytiotrophoblast into the Identified in 1983 [78], the function and physiological properties of
maternal circulation, contain high amounts of PP13 [62]. PP13 were largely unknown until recently [77]. As one of the 56 pla-
In cases with preeclampsia, the level of PP13 mRNA expression in cental proteins known today, its sequence revealed that it belongs to the
the placenta is significantly reduced. At term, there is a 3.5-fold de- family of galectins, this is why it is also termed galectin 13. Galectins
crease in PP13 mRNA in preeclampsia cases compared to controls bind to beta-galactoside-specific lectins by their evolutionarily con-
[63,64]. Interestingly, this decrease in PP13 expression is already pre- served carbohydrate-binding domains [79]. So far, PP13 has been de-
sent in the first trimester in cases subsequently developing preeclampsia scribed to bind to beta-galactoside residues of a variety of proteins and
[65]. Reduced expression of PP13 mRNA in the first trimester of thus affecting molecular recognition and apoptosis of immune cells
pregnancy can be regarded as one of the earliest signs for the devel- without a known classical receptor entity [68,80].
opment of preeclampsia. Today, it has become clear that the effects of PP13 go far beyond
Surprisingly, levels of PP13 mRNA can also be detected in maternal interactions with immune cells. In rat and rabbit models (pregnant and
peripheral blood and seem to correlate with its expression in the pla- non-pregnant), PP13 has been recognized as a strong factor inducing
centa. Again, reduced levels of PP13 mRNA in maternal blood were expansion and dilation of the uterine vasculature [81,82]. During
detected in cases with clinical symptoms of preeclampsia [66] as well as pregnancy, administration of PP13 into rats induced systemic and re-
in asymptomatic women in the first trimester who developed pre- versible hypotension, dilation of uterine veins and arteries and resulted
eclampsia later in gestation [67]. in larger pups and placentas [82]. The vasodilatory effect of PP13 was
Within the placenta proper, the PP13 protein is mostly located assessed using isolated uterine arteries from non-pregnant and pregnant
within the syncytiotrophoblast and is specifically localized at the apical rats. Although no direct receptor could be identified, relaxation of ar-
membrane of the syncytiotrophoblast [63,68–70]. Im- teries was mediated by signaling pathways of nitric oxide and pros-
munohistochemical staining of PP13 reveals highest intensities around taglandin [83].
weeks six to seven with a decreased intensity at the onset of maternal Interestingly, the dilating effect of PP13 on uterine vessels is dif-
blood flow into the placenta, i.e. weeks 12 to 14 [70]. Villous as well as ferent in veins and arteries. The larger the uterine vein (from radial via
extravillous cytotrophoblasts do not show any staining for PP13 in arcuate to main uterine vein), the larger the dilating effect. In uterine
immunohistochemistry [69,70]. arteries, the effect is in the opposite direction. The smaller the uterine
Interestingly, first trimester villous explants from healthy pregnan- artery (from main uterine via arcuate to radial artery), the larger is the
cies release twice as much PP13 as explants from term controls [71]. effect [82,84].
This feature explains why the release of PP13 from a small first tri-
mester placenta produces the same maternal serum PP13 level as a 5. New thinking: concerted action of PP13 and (anti-) angiogenic
much larger early third trimester placenta [69]. factors during preeclampsia

4.2. Use of PP13 in pregnancy Although the detailed etiology of preeclampsia is still far from being
deciphered, the intensive work on predictive biomarkers for pre-
PP13 is one of the markers with the earliest changes related to eclampsia has helped in putting some further pieces together to unravel
preeclampsia. This protein has been detected in maternal serum already how this syndrome may develop (Fig. 1).
at five weeks of gestation with significant changes monitored at week While the initial insult on the placenta is still unclear, the first effect
six in cases subsequently developing preeclampsia [69,72]. In normal that is visible today is the reduced level of PP13 in maternal blood
pregnancy, maternal serum levels of PP13 remain constant until the already at six weeks in women developing preeclampsia later in preg-
beginning of the third trimester and increase until term to reach levels nancy [69]. This may already prime the maternal vascular system and

4
B. Huppertz BBA - Molecular Basis of Disease 1866 (2020) 165349

Fig. 1. Interaction of (anti-) angiogenic factors and placental proteins in the course of the development of preeclampsia.
The left part of the scheme represents the normal release of factors during healthy pregnancy. The right part of the scheme represents the altered release of factors
during the development of preeclampsia. One of the earliest signs of developing preeclampsia is a reduced PP13 level in maternal blood (week 6), which may already
have a negative impact on the maternal vasculature and prime maternal vessels. This and the release of placental “factors” may in turn lead to endothelial activation
and thus increased levels of sFlt-1 leading to reduced levels of free PGF. This could lead to a further reduction of vessel dilation, which in the third trimester may be
counteracted by an increased release of PP13.

reduce vessel dilation necessary to adapt to the needs of pregnancy. the soluble VEGF receptor sFlt-1 can be observed in sepsis (Fig. 2),
This reduced level of PP13 very early in gestation may result in in- which has been described as a marker for the severity of sepsis [36,51].
creased arterial stiffness already at weeks 11 to 13, which leads to Increased levels of sFlt-1 in turn reduce the level of free VEGF and thus
subclinical values of increased blood pressure [85] (Fig. 1). may reduce the negative effects of this factor during sepsis.
The ongoing activation of the maternal vascular system by so far Also in preeclampsia, increased levels of sFlt-1 [37,38], paralleled
unknown placental factors may result in an increased release of VEGF, by reduced levels of free VEGF [89] have been reported at the time of
similar to what can be seen in sepsis [51] (Fig. 2). In sepsis, bacterial diagnosis in patients with preeclampsia, while the total concentration
factors induce a cascade of events including increased secretion of of VEGF does not change [40].
VEGF, which in turn activates signaling cascades that increase leuko- Sequestration of VEGF by sFlt-1 to hinder leakage, clot formation
cyte adhesion and clot formation. Respective changes of the clotting and leukocyte adhesion is achieved with the price of less dilation of
system as well as increased leukocyte adhesion have been described for vessels. This may in turn activate secretion of PP13 from the placenta in
preeclampsia as well [86,87] (Fig. 2). Spiezia et al. [87] showed an cases with preeclampsia – this time different to sepsis. Since PP13 is
increase in clot firmness as well as a decrease in blood fibrinolysis in using different signaling pathways compared to the angiogenic system,
women suffering from preeclampsia. vessel dilation by PP13 may be achieved in the presence of high sFlt-1
In addition, in sepsis activated VEGF signaling induces dilation of levels (Figs. 1, 2). Depending on the maternal system, this may or may
vessels as well as vascular leakage and formation of edema [51]. The not be sufficient to significantly reduce blood pressure. Hence, this
dilatory and leakage effects of VEGF have recently been shown outside could explain why some women with high sFlt-1 levels do not develop
of sepsis in an in vivo mouse model [88]. These authors showed a di- preeclampsia.
latory effect of VEGF on venules (diameter increase by factor of 1.7) Although the above comparison reveals a number of similarities
and a much stronger dilatory effect on capillaries (diameter increase by between sepsis and preeclampsia, there are also major differences in-
a factor of 3.0). Again, leakage and thus formation of edema are a ty- cluding blood pressure (hypotension versus hypertension) and of course
pical feature of preeclampsia and have long been part of the diagnostic bacterial infection [90] (Fig. 2).
triage with hypertension, proteinuria and edema.
Maybe as a counteracting event, increased expression and release of

5
B. Huppertz BBA - Molecular Basis of Disease 1866 (2020) 165349

Delivery-Trends.jsp?utm_source=ahrq&utm_medium=en1&utm_term=&utm_
content=1&utm_campaign=ahrq_en4_25_2017.
[2] American College of Obstetricians and Gynecologists; Task Force on Hypertension
in Pregnancy, Hypertension in pregnancy. Report of the American College of
Obstetricians and Gynecologists' Task Force on Hypertension in Pregnancy, Obstet.
Gynecol. 122 (2013) 1122–1131.
[3] D.L. Rolnik, D. Wright, L.C. Poon, N. O'Gorman, A. Syngelaki, C. de Paco Matallana,
R. Akolekar, S. Cicero, D. Janga, M. Singh, F.S. Molina, N. Persico, J.C. Jani,
W. Plasencia, G. Papaioannou, K. Tenenbaum-Gavish, H. Meiri, S. Gizurarson,
K. Maclagan, K.H. Nicolaides, Aspirin versus placebo in pregnancies at high risk for
preterm preeclampsia, N. Engl. J. Med. 377 (2017) 613–622.
[4] A. Conde-Agudelo, J. Villar, M. Lindheimer, World Health Organization systematic
review of screening tests for preeclampsia, Obstet. Gynecol. 104 (2004) 1367–1391.
[5] R. Townsend, P. O'Brien, A. Khalil, Current best practice in the management of
hypertensive disorders in pregnancy, Integr. Blood Press. Control 9 (2016) 79–94.
[6] S. Verlohren, K. Melchiorre, A. Khalil, B. Thilaganathan, Uterine artery Doppler,
birth weight and timing of onset of pre-eclampsia: providing insights into the dual
etiology of late-onset pre-eclampsia, Ultrasound Obstet. Gynecol. 44 (2014)
293–298.
[7] P. von Dadelszen, L.A. Magee, J.M. Roberts, Subclassification of preeclampsia,
Hypertens. Pregnancy 22 (2003) 143–148.
[8] N. O'Gorman, D. Wright, A. Syngelaki, R. Akolekar, A. Wright, L.C. Poon,
K.H. Nicolaides, Competing risks model in screening for preeclampsia by maternal
factors and biomarkers at 11–13 weeks gestation, Am. J. Obstet. Gynecol. 214
(2016) 103.e1–103.e12.
[9] B. Huppertz, Placental origins of preeclampsia: challenging the current hypothesis,
Hypertension 51 (2008) 970–975.
[10] B. Huppertz, The critical role of abnormal trophoblast development in the etiology
of preeclampsia, Curr. Pharm. Biotechnol. (Apr 26 2018), https://doi.org/10.2174/
1389201019666180427110547 (Epub ahead of print).
[11] H. Perry, A. Khalil, B. Thilaganathan, Preeclampsia and the cardiovascular system:
an update, Trends Cardiovasc. Med. (May 15 2018), https://doi.org/10.1016/j.tcm.
2018.04.009 (pii: S1050-1738(18)30059-8, Epub ahead of print).
[12] H. Zeisler, E. Llurba, F.J. Chantraine, M. Vatish, A.C. Staff, M. Sennström,
M. Olovsson, S.P. Brennecke, H. Stepan, D. Allegranza, M. Schoedl, S. Grill,
M. Hund, S. Verlohren, The sFlt-1/PlGF ratio: ruling out pre-eclampsia for up to 4
weeks and the value of retesting, Ultrasound Obstet. Gynecol. (Jul 16 2018),
https://doi.org/10.1002/uog.19178 (Epub ahead of print).
[13] D.E. Clark, S.K. Smith, Y. He, K.A. Day, D.R. Licence, A.N. Corps, R. Lammoglia,
D.S. Charnock-Jones, A vascular endothelial growth factor antagonist is produced
by the human placenta and released into the maternal circulation, Biol. Reprod. 59
(1998) 1540–1548.
Fig. 2. Similarities and differences between sepsis and preeclampsia. [14] D.E. Clark, S.K. Smith, D. Licence, A.L. Evans, D.S. Charnock-Jones, Comparison of
Bacterial or placental factors may have a major impact on the vascular system expression patterns for placenta growth factor, vascular endothelial growth factor
of an adult and thus lead to activation/dysregulation of endothelial cells. This (VEGF), VEGF-B and VEGF-C in the human placenta throughout gestation, J.
in turn may result in an increase of free VEGF with the respective effects of this Endocrinol. 159 (1998) 459–467.
[15] A. Farina, A. Sekizawa, P. De Sanctis, Y. Purwosunu, T. Okai, D.H. Cha, J.H. Kang,
factor. The vascular system may counteract this increased VEGF by increasing C. Vicenzi, A. Tempesta, N. Wibowo, L. Valvassori, N. Rizzo, Gene expression in
sFlt-1 to scavenge VEGF and reduce its negative effects. Besides the above si- chorionic villous samples at 11 weeks' gestation from women destined to develop
milarities between sepsis and preeclampsia, there are of course a number of preeclampsia, Prenat. Diagn. 28 (2008) 956–961.
differences shown in the scheme as well. Finally, in preeclampsia PP13 may act [16] V. Tsatsaris, F. Goffin, C. Munaut, J.F. Brichant, M.R. Pignon, A. Noel, J.P. Schaaps,
as a rescue response that is not present in sepsis. D. Cabrol, F. Frankenne, J.M. Foidart, Overexpression of the soluble vascular en-
dothelial growth factor receptor in preeclamptic patients: pathophysiological con-
sequences, J. Clin. Endocrinol. Metab. 88 (2003) 5555–5563.
[17] T. Hod, A.S. Cerdeira, S.A. Karumanchi, Molecular mechanisms of preeclampsia,
6. Conclusions
Cold Spring Harb. Perspect. Med. 5 (2015) a023473.
[18] A.C. Staff, S.J. Benton, P. von Dadelszen, J.M. Roberts, R.N. Taylor, R.W. Powers,
Preeclampsia is still the syndrome of hypotheses; however, so far D.S. Charnock-Jones, C.W. Redman, Redefining preeclampsia using placenta-de-
the different aspects of marker actions have not been combined. For the rived biomarkers, Hypertension 61 (2013) 932–942.
[19] S.E. Maynard, J.Y. Min, J. Merchan, K.H. Lim, J. Li, S. Mondal, T.A. Libermann,
first time, the effects of various predictive biomarkers for preeclampsia J.P. Morgan, F.W. Sellke, I.E. Stillman, F.H. Epstein, V.P. Sukhatme,
have been assembled into a logical order to identify causes and con- S.A. Karumanchi, Excess placental soluble fms-like tyrosine kinase 1 (sFlt1) may
sequences. This combination revealed that there may well be a direct contribute to endothelial dysfunction, hypertension, and proteinuria in pre-
eclampsia, J. Clin. Invest. 111 (2003) 649–658.
interaction between placental proteins and angiogenic factors and their [20] C.W. Redman, I.L. Sargent, Placental stress and pre-eclampsia: a revised view,
receptors. Recent insight into the effects of PP13 allowed a better un- Placenta 30 (Suppl. A) (2009) S38–S42.
derstanding of the interactions between angiogenic factors/receptors [21] E. Lecarpentier, A. Atallah, J. Guibourdenche, M. Hebert-Schuster, S. Vieillefosse,
A. Chissey, B. Haddad, G. Pidoux, D. Evain-Brion, A. Barakat, T. Fournier,
and placental proteins. This makes it tempting to speculate on a new V. Tsatsaris, Fluid shear stress promotes placental growth factor upregulation in
therapeutic option where replenishment of PP13 early in pregnancy human syncytiotrophoblast through the cAMP-PKA signaling pathway,
may help in preventing or reducing severity of preeclampsia. Hypertension 68 (2016) 1438–1446.
[22] L. Ehrlich, A. Hoeller, M. Golic, F. Herse, F.H. Perschel, W. Henrich, R. Dechend,
B. Huppertz, S. Verlohren, Increased placental sFlt-1 but unchanged PlGF expres-
Transparency document sion in late-onset preeclampsia, Hypertens. Pregnancy 36 (2017) 175–185.
[23] A. Hoeller, L. Ehrlich, M. Golic, F. Herse, F.H. Perschel, M. Siwetz, W. Henrich,
R. Dechend, B. Huppertz, S. Verlohren, Placental expression of sFlt-1 and PlGF in
The Transparency document associated with this article can be
early preeclampsia vs. early IUGR vs. age-matched healthy pregnancies, Hypertens.
found, in online version. Pregnancy 36 (2017) 151–160.
[24] A. Sekizawa, Y. Purwosunu, A. Farina, H. Shimizu, M. Nakamura, N. Wibowo,
References N. Rizzo, T. Okai, Prediction of pre-eclampsia by an analysis of placenta-derived
cellular mRNA in the blood of pregnant women at 15–20 weeks of gestation, Br. J.
Obstet. Gynaecol. 110 (2010) 557–564.
[1] K.R. Fingar, I. Mabry-Hernandez, Q. Ngo-Metzger, T. Wolff, C.A. Steiner, [25] B. Huppertz, An updated view on the origin and use of angiogenic biomarkers for
A. Elixhauser, Statistical Brief #222. Healthcare Cost and Utilization Project preeclampsia, Expert. Rev. Mol. Diagn. (2018 Nov 9) 1–9, https://doi.org/10.1080/
(HCUP), Agency for Healthcare Research and Quality, Rockville, MD, April 14737159.2018.1546579.
2017www.hcup-us.ahrq.gov/reports/statbriefs/sb222-Preeclampsia-Eclampsia- [26] K. Benirschke, P. Kaufmann, R.N. Baergen, Pathology of the Human Placenta, 5th

6
B. Huppertz BBA - Molecular Basis of Disease 1866 (2020) 165349

edition, Springer, New York, 2006. O. Feron, Delivery of soluble VEGF receptor 1 (sFlt1) by gene electrotransfer as a
[27] E. Bianconi, A. Piovesan, F. Facchin, A. Beraudi, R. Casadei, F. Frabetti, L. Vitale, new antiangiogenic cancer therapy, Mol. Pharm. 8 (2011) 701–708.
M.C. Pelleri, S. Tassani, F. Piva, S. Perez-Amodio, P. Strippoli, S. Canaider, An es- [51] N.I. Shapiro, P. Schuetz, K. Yano, M. Sorasaki, S.M. Parikh, A.E. Jones, S. Trzeciak,
timation of the number of cells in the human body, Ann. Hum. Biol. 40 (2013) L. Ngo, W.C. Aird, The association of endothelial cell signaling, severity of illness,
463–471. and organ dysfunction in sepsis, Crit. Care 14 (2010) R182.
[28] S. Zamudio, M. Borges, L. Echalar, O. Kovalenko, E. Vargas, T. Torricos, A.A. Khan, [52] P. Dumnicka, M. Sporek, M. Mazur-Laskowska, P. Ceranowicz, M. Kuźniewski,
M. Alvarez, N.P. Illsley, Maternal and fetoplacental hypoxia do not alter circulating R. Drożdż, T. Ambroży, R. Olszanecki, B. Kuśnierz-Cabala, Serum soluble Fms-like
angiogenic growth effectors during human pregnancy, Biol. Reprod. 90 (2014) 1–9. tyrosine kinase 1 (sFlt-1) predicts the severity of acute pancreatitis, Int. J. Mol. Sci.
[29] M. Autiero, A. Luttun, M. Tjwa, P. Carmeliet, Placental growth factor and its re- 17 (2016) 2038.
ceptor, vascular endothelial growth factor receptor-1: novel targets for stimulation [53] D. Chelli, A. Hamdi, S. Saoudi, A.A. Jenayah, A. Zagre, H. Jguerim, C. Bedis, E. Sfar,
of ischemic tissue revascularization and inhibition of angiogenic and inflammatory Clinical assessment of soluble FMS-like tyrosine kinase-1/placental growth factor
disorders, J. Thromb. Haemost. 1 (2003) 1356–1370. ratio for the diagnostic and the prognosis of preeclampsia in the second trimester,
[30] C.J. Green, P. Lichtlen, N.T. Huynh, M. Yanovsky, K.R. Laderoute, W. Schaffner, Clin. Lab. 62 (2016) 1927–1932.
B.J. Murphy, Placenta growth factor gene expression is induced by hypoxia in fi- [54] M. Medinger, J. Halter, D. Heim, A. Buser, S. Gerull, M. Stern, J. Passweg,
broblasts: a central role for metal transcription factor-1, Cancer Res. 61 (2001) Angiogenic markers in plasma cell myeloma patients treated with novel agents,
2696–2703. Anticancer Res. 35 (2015) 1085–1090.
[31] F. Cianfarani, G. Zambruno, L. Brogelli, F. Sera, P.M. Lacal, M. Pesce, [55] M. Ziche, D. Maglione, D. Ribatti, L. Morbidelli, C.T. Lago, M. Battisti, I. Paoletti,
M.C. Capogrossi, C.M. Failla, M. Napolitano, T. Odorisio, Placenta growth factor in A. Barra, M. Tucci, G. Parise, V. Vincenti, H.J. Granger, G. Viglietto, M.G. Persico,
diabetic wound healing: altered expression and therapeutic potential, Am. J. Pathol. Placenta growth factor-1 is chemotactic, mitogenic, and angiogenic, Lab. Investig.
169 (2006) 1167–1182. 76 (1997) 517–531.
[32] H.P. Gerber, T.H. Vu, A.M. Ryan, J. Kowalski, Z. Werb, N. Ferrara, VEGF couples [56] M. Dewerchin, P. Carmeliet, PlGF: a multitasking cytokine with disease-restricted
hypertrophic cartilage remodeling, ossification and angiogenesis during en- activity, Cold Spring Harb. Perspect. Med. 2 (2012) a011056.
dochondral bone formation, Nat. Med. 5 (1999) 623–628. [57] P. Carmeliet, L. Moons, A. Luttun, V. Vincenti, V. Compernolle, M. DeMol, Y. Wu,
[33] B. Barleon, S. Sozzani, D. Zhou, H.A. Weich, A. Mantovani, D. Marmé, Migration of F. Bono, L. Devy, H. Beck, D. Scholz, T. Acker, T. DiPalma, M. Dewerchin, A. Noel,
human monocytes in response to vascular endothelial growth factor (VEGF) is I. Stalmans, A. Barra, S. Blacher, T. VandenDriessche, A. Ponten, U. Eriksson,
mediated via the VEGF receptor flt-1, Blood 87 (1996) 3336–3343. K.H. Plate, J.M. Foidart, W. Schaper, D.S. Charnock-Jones, D.J. Hicklin,
[34] A. Ishida, J. Murray, Y. Saito, C. Kanthou, O. Benzakour, M. Shibuya, E.S. Wijelath, J.M. Herbert, D. Collen, M.G. Persico, Synergism between vascular endothelial
Expression of vascular endothelial growth factor receptors in smooth muscle cells, growth factor and placental growth factor contributes to angiogenesis and plasma
J. Cell. Physiol. 188 (2001) 359–368. extravasation in pathological conditions, Nat. Med. 7 (2001) 575–583.
[35] J.S. Krüssel, E.M. Casañ, F. Raga, J. Hirchenhain, Y. Wen, H.Y. Huang, P. Bielfeld, [58] G. Osol, G. Celia, N. Gokina, C. Barron, E. Chien, M. Mandala, L. Luksha,
M.L. Polan, Expression of mRNA for vascular endothelial growth factor trans- K. Kublickiene, Placental growth factor (PlGF) is a potent vasodilator of rat and
membraneous receptors Flt1 and KDR, and the soluble receptor sflt in cycling human resistance arteries, Am. J. Physiol. Heart Circ. Physiol. 294 (2008)
human endometrium, Mol. Hum. Reprod. 5 (1999) 452–458. H1381–H1387.
[36] C.M. Failla, M. Carbo, V. Morea, Positive and negative regulation of angiogenesis by [59] J. Gavard, J.S. Gutkind, VEGF controls endothelial-cell permeability by promoting
soluble vascular endothelial growth factor receptor-1, Int. J. Mol. Sci. 19 (2018) the β-arrestin-dependent endocytosis of VE-cadherin, Nat. Cell Biol. 8 (2006)
1306–1322. 1223–1234.
[37] R.W. Powers, J.M. Roberts, K.M. Cooper, M.J. Gallaher, M.P. Frank, G.F. Harger, [60] Y.L. Hashambhoy, J.C. Chappell, S.M. Peirce, V.L. Bautch, F. MacGabhann,
R.B. Ness, Maternal serum soluble fms-like tyrosine kinase 1 concentrations are not Computational modeling of interacting VEGF and soluble VEGF receptor con-
increased in early pregnancy and decrease more slowly postpartum in women who centration gradients, Front. Physiol. 2 (2011) 1–12.
develop preeclampsia, Am. J. Obstet. Gynecol. 193 (2005) 185–191. [61] N.G. Than, R. Romero, M. Goodman, A. Weckle, J. Xing, Z. Dong, Y. Xu, F. Tarquini,
[38] S. Rana, S.A. Karumanchi, R.J. Levine, S. Venkatesha, J.A. Rauh-Hain, H. Tamez, A. Szilagyi, P. Gal, Z. Hou, A.L. Tarca, C.J. Kim, J.S. Kim, S. Haidarian, M. Uddin,
R. Thadhani, Sequential changes in antiangiogenic factors in early pregnancy and H. Bohn, K. Benirschke, J. Santolaya-Forgas, L.I. Grossman, O. Erez, S.S. Hassan,
risk of developing preeclampsia, Hypertension 50 (2007) 137–142. P. Zavodszky, Z. Papp, D.E. Wildman, A primate subfamily of galectins expressed at
[39] R. Romero, J.K. Nien, J. Espinoza, D. Todem, W. Fu, H. Chung, J.P. Kusanovic, the maternal-fetal interface that promote immune cell death, Proc. Natl. Acad. Sci.
F. Gotsch, O. Erez, S. Mazaki-Tovi, R. Gomez, S. Edwin, T. Chaiworapongsa, U. S. A. 106 (2009) 9731–9736.
R.J. Levine, S.A. Karumanchi, A longitudinal study of angiogenic (placental growth [62] M. Sammar, R. Dragovic, H. Meiri, M. Vatish, A. Sharabi-Nov, I. Sargent,
factor) and anti-angiogenic (soluble endoglin and soluble vascular endothelial C. Redman, D. Tannetta, Reduced placental protein 13 (PP13) in placental derived
growth factor receptor-1) factors in normal pregnancy and patients destined to syncytiotrophoblast extracellular vesicles in preeclampsia - a novel tool to study the
develop preeclampsia and deliver a small for gestational age neonate, J. Matern. impaired cargo transmission of the placenta to the maternal organs, Placenta 66
Fetal Neonatal Med. 21 (2008) 9–23. (2018) 17–25.
[40] A. Virtanen, O. Huttala, K. Tihtonen, T. Toimela, T. Heinonen, J. Uotila, Angiogenic [63] N.G. Than, O. Abdul Rahman, R. Magenheim, B. Nagy, T. Fule, B. Hargitai,
capacity in pre-eclampsia and uncomplicated pregnancy estimated by assay of an- M. Sammar, P. Hupuczi, A.L. Tarca, G. Szabo, I. Kovalszky, H. Meiri, I. Sziller,
giogenic proteins and an in vitro vasculogenesis/angiogenesis test, Angiogenesis J. Rigo Jr., R. Romero, Z. Papp, Placental protein 13 (galectin-13) has decreased
(Jul 12 2018), https://doi.org/10.1007/s10456-018-9637-2 (Epub ahead of print). placental expression but increased shedding and maternal serum concentrations in
[41] E. Litwińska, M. Litwińska, P. Oszukowski, K. Szaflik, P. Kaczmarek, Combined patients presenting with preterm pre-eclampsia and HELLP syndrome, Virchows
screening for early and late pre-eclampsia and intrauterine growth restriction by Arch. 453 (2008) 387–400.
maternal history, uterine artery Doppler, mean arterial pressure and biochemical [64] M. Sammar, S. Nisemblat, Z. Fleischfarb, A. Golan, O. Sadan, H. Meiri, B. Huppertz,
markers, Adv. Clin. Exp. Med. 26 (2017) 439–448. R. Gonen, Placenta-bound and body fluid PP13 and its mRNA in normal pregnancy
[42] L.C. Poon, C. Lesmes, D.M. Gallo, R. Akolekar, K.H. Nicolaides, Prediction of small- compared to preeclampsia, HELLP and preterm delivery, Placenta 32 (2011)
for-gestational-age neonates: screening by biophysical and biochemical markers at S30–S36.
19–24 weeks, Ultrasound Obstet. Gynecol. 46 (2015) 437–445. [65] A. Sekizawa, Y. Purwosunu, S. Yoshimura, M. Nakamura, H. Shimizu, T. Okai,
[43] C. Lesmes, D.M. Gallo, R. Gonzalez, L.C. Poon, K.H. Nicolaides, Prediction of small- N. Rizzo, A. Farina, PP13 mRNA expression in trophoblasts from preeclamptic
for-gestational-age neonates: screening by maternal serum biochemical markers at placentas, Reprod. Sci. 16 (2009) 408–413.
19–24 weeks, Ultrasound Obstet. Gynecol. 46 (2015) 341–349. [66] H. Shimizu, A. Sekizawa, Y. Purwosunu, M. Nakamura, A. Farina, N. Rizzo, T. Okai,
[44] U.V. Ukah, J.A. Hutcheon, B. Payne, M.D. Haslam, M. Vatish, J.M. Ansermino, PP13 mRNA expression in the cellular component of maternal blood as a marker for
H. Brown, L.A. Magee, P. von Dadelszen, Placental growth factor as a prognostic preeclampsia, Prenat. Diagn. 29 (2009) 1231–1236.
tool in women with hypertensive disorders of pregnancy: a systematic review, [67] A. Farina, C. Zucchini, A. Sekizawa, Y. Purwosunu, P. de Sanctis, G. Santarsiero,
Hypertension 70 (2017) 1228–1237. N. Rizzo, D. Morano, T. Okai, Performance of messenger RNAs circulating in ma-
[45] S.C. Wei, P.N. Tsao, S.C. Yu, C.T. Shun, J.J. Tsai-Wu, C.H. Wu, Y.N. Su, F.J. Hsieh, ternal blood in the prediction of preeclampsia at 10–14 weeks, Am. J. Obstet.
J.M. Wong, Placenta growth factor expression is correlated with survival of patients Gynecol. 203 (2010) 575.e1–575.e7.
with colorectal cancer, Gut 54 (2005) 666–672. [68] N.G. Than, E. Pick, S. Bellyei, A. Szigeti, O. Burger, Z. Berente, T. Janaky,
[46] H.M.C. Shantha Kumara, J.C. Cabot, X. Yan, S.A. Herath, M. Luchtefeld, A. Boronkai, H. Kliman, H. Meiri, H. Bohn, G.N. Than, B. Sumegi, Functional
M.F. Kalady, D.L. Feingold, R. Baxter, R.L. Whelan, Minimally invasive colon re- analyses of placental protein 13/galectin-13, Eur. J. Biochem. 271 (2004)
section is associated with a persistent increase in plasma PlGF levels following 1065–1078.
cancer resection, Surg. Endosc. 25 (2011) 2153–2158. [69] B. Huppertz, M. Sammar, I. Chefetz, P. Neumaier-Wagner, C. Bartz, H. Meiri,
[47] Q. Meng, P. Duan, L. Li, Y. Miao, Expression of placenta growth factor is associated Longitudinal determination of serum placental protein 13 during development of
with unfavorable prognosis of advanced-stage serous ovarian cancer, Tohoku J. preeclampsia, Fetal Diagn. Ther. 24 (2008) 230–236.
Exp. Med. 244 (2018) 291–296. [70] H.J. Kliman, M. Sammar, Y.I. Grimpel, S.K. Lynch, K.M. Milano, E. Pick, J. Bejar,
[48] M. Toi, H. Bando, T. Ogawa, M. Muta, C. Hornig, H.A. Weich, Significance of A. Arad, J.J. Lee, H. Meiri, R. Gonen, Placental protein 13 and decidual zones of
vascular endothelial growth factor (VEGF)/soluble VEGF receptor-1 relationship in necrosis: an immunologic diversion that may be linked to preeclampsia, Reprod.
breast cancer, Int. J. Cancer 9 (2002) 14–18. Sci. 19 (2012) 16–30.
[49] K.M. Scheufler, J. Drevs, V. van Velthoven, P. Reusch, J. Klisch, H.G. Augustin, [71] Y.I. Grimpel, V. Kivity, A. Cohen, H. Meiri, M. Sammar, R. Gonen, B. Huppertz,
J. Zentner, D. Marme, Implications of vascular endothelial growth factor, sFlt-1, and Effects of calcium, magnesium, low-dose aspirin and low-molecular-weight heparin
sTie-2 in plasma, serum and cerebrospinal fluid during cerebral ischemia in man, J. on the release of PP13 from placental explants, Placenta 32 (2011) S55–S64.
Cereb. Blood Flow Metab. 23 (2003) 99–110. [72] R. Gonen, R. Shahar, Y.I. Grimpel, I. Chefetz, M. Sammar, H. Meiri, Y. Gibor,
[50] J. Verrax, F. Defresne, F. Lair, G. Vandermeulen, G. Rath, C. Dessy, V. Préat, Placental protein 13 as an early marker for pre-eclampsia: a prospective

7
B. Huppertz BBA - Molecular Basis of Disease 1866 (2020) 165349

longitudinal study, Br. J. Obstet. Gynaecol. 115 (2008) 1465–1472. [82] S. Gizurarson, E.R. Sigurdardottir, H. Meiri, B. Huppertz, M. Sammar, A. Sharabi-
[73] I. Chafetz, I. Kuhnreich, M. Sammar, Y. Tal, Y. Giber, H. Meiri, H. Cuckle, M. Wolf, Nov, M. Mandalá, G. Osol, Placental protein 13 administration to pregnant rats
First-trimester placental protein 13 screening for preeclampsia and intrauterine lowers blood pressure and augments fetal growth and venous remodeling, Fetal
growth restriction, Am. J. Obstet. Gynecol. 197 (2007) 35.e1–35.e7. Diagn. Ther. 39 (2016) 56–63.
[74] R. Romero, J.P. Kusanovic, N.G. Than, O. Erez, F. Gotsch, J. Espinoza, S. Edwin, [83] T. Drobnjak, S. Gizurarson, N.I. Gokina, H. Meiri, M. Mandalá, B. Huppertz, G. Osol,
I. Chefetz, R. Gomez, J.K. Nien, M. Sammar, B. Pineles, S.S. Hassan, H. Meiri, Y. Tal, Placental protein 13 (PP13)-induced vasodilation of resistance arteries from preg-
I. Kuhnreich, Z. Papp, H.S. Cuckle, First-trimester maternal serum PP13 in the risk nant and nonpregnant rats occurs via endothelial-signaling pathways, Hypertens.
assessment for preeclampsia, Am. J. Obstet. Gynecol. 199 (2008) 122.e1–122.e11. Pregnancy 36 (2017) 186–195.
[75] A. Khalil, N.J. Cowans, K. Spencer, S. Goichman, H. Meiri, K. Harrington, First- [84] T. Drobnjak, A.M. Jónsdóttir, H. Helgadóttir, M.S. Runólfsdóttir, H. Meiri,
trimester markers for the prediction of pre-eclampsia in women with a-priori high M. Sammar, G. Osol, M. Mandalá, B. Huppertz, S. Gizurarson, Placental protein 13
risk, Ultrasound Obstet. Gynecol. 35 (2010) 671–679. (PP13) stimulates rat uterine vessels after slow subcutaneous administration, Drug
[76] H. Meiri, M. Sammar, A. Herzog, Y.I. Grimpel, G. Fihaman, A. Cohen, V. Kivity, Des. Devel. Ther. (2018) (under review).
A. Sharabi-Nov, R. Gonen, Prediction of preeclampsia by placental protein 13 and [85] A. Khalil, R. Akolekar, A. Syngelak, M. Elkhouli, K.H. Nicolaides, Maternal hemo-
background risk factors and its prevention by aspirin, J. Perinat. Med. 42 (2014) dynamics at 11–13 weeks' gestation and risk of pre-eclampsia, Ultrasound Obstet.
591–601. Gynecol. 40 (2012) 28–34.
[77] B. Huppertz, H. Meiri, S. Gizurarson, G. Osol, M. Sammar, Placental protein 13 [86] A.M. Gil-Villa, L.V. Norling, C.N. Serhan, D. Cordero, M. Rojas, A. Cadavid, Aspirin
(PP13): a new biological target shifting individualized risk assessment to persona- triggered-lipoxin A4 reduces the adhesion of human polymorphonuclear neu-
lized drug design combating pre-eclampsia, Hum. Reprod. Update 19 (2013) trophils to endothelial cells initiated by preeclamptic plasma, Prostaglandins
391–405. Leukot. Essent. Fat. Acids 87 (2012) 127–134.
[78] H. Bohn, W. Kraus, W. Winckler, Purification and characterization of two new so- [87] L. Spiezia, G. Bogana, E. Campello, S. Maggiolo, E. Pelizzaro, C.D. Carbonare,
luble placental tissue proteins (PP13 and PP17), Oncodev. Biol. Med. 4 (1983) M.T. Gervasi, P. Simioni, Whole blood thromboelastometry profiles in women with
343–350. preeclampsia, Clin. Chem. Lab. Med. 53 (2015) 1793–1798.
[79] N.G. Than, R. Romero, C.J. Kim, M.R. McGowen, Z. Papp, D.E. Wildman, Galectins: [88] N. Honkura, M. Richards, B. Laviña, M. Sáinz-Jaspeado, C. Betsholtz, L. Claesson-
guardians of eutherian pregnancy at the maternal-fetal interface, Trends Welsh, Intravital imaging-based analysis tools for vessel identification and assess-
Endocrinol. Metab. 23 (2012) 23–31. ment of concurrent dynamic vascular events, Nat. Commun. 9 (2018) 2746.
[80] B. Visegrady, N.G. Than, F. Kilar, B. Sümegi, G.N. Than, H. Bohn, Homology [89] K. Adu-Bonsaffoh, D.A. Antwi, B. Gyan, S.A. Obed, Endothelial dysfunction in the
modelling and molecular dynamics studies of human placental tissue protein 13 pathogenesis of pre-eclampsia in Ghanaian women, BMC Physiol. 17 (2017) 5.
(galectin-13), Protein Eng. 14 (2001) 875–880. [90] M. Bonet, J.P. Souza, E. Abalos, B. Fawole, M. Knight, S. Kouanda, P. Lumbiganon,
[81] S. Gizurarson, B. Huppertz, G. Osol, J.O. Skarphedinsson, M. Mandala, H. Meiri, A. Nabhan, R. Nadisauskiene, V. Brizuela, A.M. Gülmezoglu, The global maternal
Effects of placental protein 13 on the cardiovascular system in gravid and non- sepsis study and awareness campaign (GLOSS): study protocol, Reprod. Health 15
gravid rodents, Fetal Diagn. Ther. 33 (2013) 257–264. (2018) 16.

You might also like