Hostname: page-component-6bf8c574d5-gr6zb Total loading time: 0 Render date: 2025-02-14T22:36:11.604Z Has data issue: false hasContentIssue false

Receptors for growth and development of Schistosoma mansoni

Published online by Cambridge University Press:  14 February 2025

Iman F. Abou-El-Naga*
Affiliation:
Medical Parasitology Department, Faculty of Medicine, Alexandria University, Egypt
*
Corresponding author: Iman F. Abou-El-Naga; Email: eman.abuelnaga@alexmed.edu.eg
Rights & Permissions [Opens in a new window]

Abstract

The growth and development of schistosomes are tightly regulated by various receptors throughout their life cycle. Each stage of the parasite inhabits a distinct habitat and responds to different factors that drive its growth and development. With two hosts involved in its life cycle (mammalian and snail), the parasite must go through additional free-living stages to transition between them. Moreover, communication between male and female worms is essential for the maturation of females. The ability of adult schistosomes to survive in human hosts for up to thirty years demonstrates their capacity to efficiently utilize host nutrients for metabolic processes and growth. In Schistosoma mansoni, receptors mediate the utilization of growth factors derived from both the parasite itself and the host. Nuclear receptors, in particular, collaborate with other proteins to regulate the expression of genes essential for various developmental functions. Receptors also play a pivotal role in RNA export, which is crucial for the parasite development. Additionally, neurotransmitter receptors are essential for the growth and development of larval stages. This review aims to elucidate the mechanisms by which these receptors regulate cell proliferation, differentiation, and maturation throughout the parasite life cycle. Understanding these processes could provide insights into the role of receptors in Schistosoma mansoni development and potentially lead to innovative therapeutic strategies to combat human schistosomiasis.

Type
Review Article
Copyright
© The Author(s), 2025. Published by Cambridge University Press

Introduction

Schistosomiasis is a major neglected tropical disease affecting approximately 240 million people (World Health Organization 2024). Schistosoma mansoni (S. mansoni), S. japonicum, and S. haematobium are the three main schistosome species that cause schistosomiasis in humans. S. mansoni is the causative agent of intestinal schistosomiasis in humans. The disease is usually linked with impoverished socioeconomic conditions (Abou-El-Naga Reference Abou-El-Naga2015) and is characterized by hepatosplenomegaly, portal hypertension, anemia, and eosinophilia (el Zawawy et al. Reference el Zawawy, el Nassery, al Azzouni, Abou el Naga, el Temsahi and Awadalla1995; Shaker et al. Reference Shaker, Samy and Ashour2014). There is currently no effective vaccine against the parasite. Various molluscicides have been evaluated to control the intermediate snail host of the disease (Younis et al. Reference Younis, Abou-El-Naga and Radwan2023; Zheng et al. Reference Zheng, Deng, Zhong, Wang, Guo and Fan2021). The primary strategy for controlling schistosomiasis is the mass drug administration of praziquantel (PZQ) (Abou-El-Naga Reference Abou-El-Naga2018). However, PZQ is ineffective against immature worms and offers no protection against re-infection (Mogahed et al. Reference Mogahed, El-Temsahy, Abou-El-Naga, Makled, Sheta and Ibrahim2023). Reports of isolates with reduced susceptibility to PZQ and the possibility of experimentally producing praziquantel resistance (Amer et al. Reference Amer, Abou-El-Naga, Boulos, Ramadan and Younis2022) highlight the risks of relying on a single therapeutic agent for a disease of this magnitude.

Adult schistosomes parasitize humans and lay eggs, many of which are eventually expelled from their definitive hosts with feces. However, some eggs fail to undergo the extravasation process needed for expulsion; instead, they are carried by the bloodstream and become trapped in the liver (Walker Reference Walker2011). In fresh waters, each egg hatches into a ciliated miracidium that infects a snail intermediate host of genus Biomphalaria.

Snails show varying degrees of susceptibility and maintain a complex relationship with the parasite (Abou-El-Naga and Radwan Reference Abou-El-Naga and Radwan2012; Abou-El-Naga et al. Reference Abou-El-Naga, Sadaka, Amer, Diab and Khedr2015; El Naga et al. Reference El Naga, Eissa, Mossallam and El-Halim2010). Inside the snail, the miracidium undergoes a dramatic transformation into an obligate asexually reproducing mother sporocyst. The proliferation of stem cells in a mother sporocyst gives rise to a new asexual stage, the daughter sporocyst. Germinal cells proliferate in a mother sporocyst to produce a daughter sporocyst. The cercariae that emerge from the daughter sporocyst are released into the water (Walker Reference Walker2011). They penetrate human skin with the aid of the fatty acids present in the skin (Hammouda et al. Reference Hammouda, Abou el Naga, el Temsahi and Sharaf1994; Salter et al. Reference Salter, Lim, Hansell, Hsieh and McKerrow2000). The cercariae transform into schistosomula that migrate into the branches of the hepatic portal vein. They ingest blood cells and grow into juvenile schistosomes in the liver. The juveniles then couple and mature into adult male and female worms, which migrate to the mesenteric veins to mate and lay eggs (Walker Reference Walker2011).

Throughout its life cycle, Schistosoma inhabits distinct habitats and alternates between mammalian and snail hosts. This dual-host life cycle requires additional free-living stages that facilitate these transitions. Each stage of the parasite inhabits a distinct habitat and responds to different factors that promote its growth and development. Furthermore, the male Schistosoma worm controls the development of the female. The sexual development of the female is determined by mating with a male schistosome worm (Severinghaus Reference Severinghaus1928). Maintaining the female mature reproductive state requires perpetual mating with a male worm, not sperm transfer (Popiel et al. Reference Popiel, Cioli and Erasmus1984). Chen et al. (Reference Chen, Wang, Gradinaru, Vu, Geboers, Naidoo, Ready, Williams, DeBerardinis, Ross and Collins2022) identified a non-ribosomal peptide synthetase that is activated in male worms when they mate with a female and determined that it is crucial for male worms to promote female development. Adult schistosomes can survive in human hosts for up to thirty years (von Lichtenberg Reference von Lichtenberg, Rollinson and Simpson1987), demonstrating their ability to effectively utilize the host’s nutrients for metabolic processes and growth (You et al. Reference You, Gobert, Jones, Zhang and McManus2011). Among trematodes, schistosomes are unique in that they have separate sexes. Adult male and female schistosomes live constantly paired, which is essential for the development of the female gonads. Females without pairing experience are sexually immature. When pairing occurs, differentiation processes are triggered that lead to maturation of the ovary and vitellarium, resulting in a sexually mature female. Unlike females, pairing-inexperienced males already have testes with differentiated spermatocytes and show no morphological differences compared to pairing-experienced males. However, pairing also induces changes in male gene expression (Lu et al. Reference Lu, Sessler, Holroyd, Hahnel, Quack, Berriman and Grevelding2016).

S. mansoni utilizes a variety of receptors to regulate its growth and development. This review seeks to clarify how these receptors regulate cell proliferation, differentiation, and maturation throughout the parasite life cycle. The findings presented here will provide a comprehensive understanding of the crucial role of these receptors in the development of S. mansoni.

With the increasing availability of genomic data from S. mansoni, an increase in studies focusing on elucidating the role of receptors in host-parasite interactions is expected. Understanding the molecular basis of receptor functions and the development of more specific receptor agonists and antagonists represents a substantial challenge for future research.

Growth factor receptors (Table 1 )

Growth factors are a group of polypeptides that play a crucial role in regulating a variety of cellular processes, including cell growth, proliferation, and differentiation. They are essential for the normal development and function of tissues and organs. Growth factors are considered a subset of cytokines. While all cytokines influence signal transduction pathways, only those cytokines affecting cell growth/differentiation signalling pathways are considered growth factors. Thus, growth factors have a positive effect on cell division, while cytokine is a neutral term in relation to whether a molecule affects proliferation. They are produced by various cell types and typically act locally in an autocrine or paracrine manner. They can circulate in the plasma and bind to specific proteins. In this bound form, they remain inactive but can be activated locally. The most important growth factors are epidermal growth factor, insulin-like growth factor, fibroblast growth factor, and transforming growth factor-ß (Stone et al. Reference Stone, Leavitt and Varacallo2023). Most growth factor receptors have tyrosine kinase activity by phosphorylating downstream protein tyrosine residues. The surface receptors for the TGF-β are an exception. When activated by the binding of TGF-β cytokines, this receptor can phosphorylate downstream proteins on serine and threonine residues (Saito et al. Reference Saito, Horie and Nagase2018). Molecular data have identified schistosome growth factor receptors (Collins et al. Reference Collins, Wang, Lambrus, Tharp, Iyer and Newmark2013; Du et al. Reference Du, McManus, French, Collinson, Sivakumaran, MacGregor, Fogarty, Jones and You2023; Wang et al. Reference Wang, Collins and Newmark2013).

Table 1. Growth factor receptors

Gene ID is extracted from the WormBase ParaSite using the reference genome for S. mansoni, SM_V10 (WormBase ParaSite 2024).

The receptor tyrosine kinases (RTKs) of S. mansoni include those receptors responsible for growth, which are four members of the epidermal growth factor receptors (EGFRs) family, two of the insulin receptor family (IRs), and two members of the fibroblast growth factor receptor (FGFRs) (Andrade et al. Reference Andrade, Nahum, Avelar, Silva, Zerlotini, Ruiz and Oliveira2011; Avelar et al. Reference Avelar, Nahu, Andrade and Oliveira2011). In addition, the schistosome genome encodes two Venus kinase receptors (VKRs), which belong to a family of RTKs originally discovered in S. mansoni (Vicogne et al. Reference Vicogne, Pin, Lardans, Capron, Noël and Dissous2003).

Epidermal growth factor receptors (Smp_165470; Smp_093930; Smp_152680; Smp_344500)

Epidermal growth factor receptors (EGFRs) are transmembrane glycoprotein and belong to the receptor tyrosine kinases (RTKs) (Grapa et al. Reference Grapa, Mocan, Gonciar, Zdrehus, Mosteanu, Pop and Mocan2019). The human EGFR is associated with the pathogenesis and progression of various types of carcinomas. In urothelial carcinoma associated with Schistosoma infection, a higher level of EGFR is found than in urothelial carcinoma of other causes (AlHariry et al. Reference AlHariry, El Saftawy, Aboulhoda, Abozamel, Alghamdi, Hamoud and Khalil Ghanam2024). In S. mansoni, EGFR homologs are predominantly expressed in the muscle of adult male and female worms, indicating that this receptor may play a role in muscle development (Ramachandran et al. Reference Ramachandran, Skelly and Shoemaker1996). Moreover, the EGFR substrate is expressed in the vitellarium and the ovary of the adult female and in the testes of the adult male worms, suggesting that this receptor may have additional functions in the gonads (Buro et al. Reference Buro, Burmeister, Quack and Grevelding2017). Maharjan et al. (Reference Maharjan, Kirk, Lawton and Walker2023) demonstrated the anterior localization of EGFR in schistosomula.

EGFR contains a conserved intracellular tyrosine kinase domain, a unique transmembrane hydrophobic domain, and an extracellular domain for binding EGF ligands. Human EGF induces EGFR autophosphorylation in adult worms and larvae of S. mansoni and increases protein and DNA synthesis in adult worms, suggesting that host hormones are involved in the regulation of schistosome development. Schistosome EGFR can bind human EGF with the same affinity as human EGFR (Ramachandran et al. Reference Ramachandran, Skelly and Shoemaker1996).

Human growth factors can modulate schistosome-signalling processes such as protein kinase C (PKC) and extracellular signal-regulated kinase (ERK) (Ressurreição et al. Reference Ressurreição, Elbeyioglu, Kirk, Rollinson, Emery, Page and Walker2016; Vicogne et al. Reference Vicogne, Cailliau, Tulasne, Browaeys, Yan, Fafeur, Vilain, Legrand, Trolet and Dissous2004). Human EGF, insulin, and insulin-like growth factor 1 were found to activate PKC and ERK at the schistosomula surface. The stimulation of these signalling by human growth factors is crucial during early host invasion, as the parasite encounters human growth factors for the first time and must rapidly adapt to the host. Host-mediated ERK activation can drive tegument remodeling, ensuring parasite survival while promoting cell growth and differentiation. Depleting of cholesterol from tegument lipid rafts, which are crucial for S. mansoni biology, disrupts EGFR/IR binding on the schistosomula surface and alters several protein kinases signalling pathways within the parasite (Ressurreição et al. Reference Ressurreição, Elbeyioglu, Kirk, Rollinson, Emery, Page and Walker2016).

Insulin receptors (IRs)

During growth and reproduction, schistosomes consume substantial amounts of energy derived primarily from the host’s nutrition. Adult S. mansoni worms absorb large amounts of blood glucose, equivalent to their dry weight every five hours, from the portal and mesenteric veins of the host (Bueding Reference Bueding1950). Glucose uptake occurs primarily through facilitated diffusion across the worm tegument. Human insulin has been shown to enhance glucose uptake in schistosomes (Ahier et al. Reference Ahier, Khayath, Vicogne and Dissous2008). Two glucose transporters, GTP1 and GTP4, play a crucial role in this process and are distributed asymmetrically on the tegument. GTP1 is located within the basal membrane and transports glucose into the underlying tissues, whereas GTP4 is located in the apical membrane. GTP4 is expressed on the parasite surface concurrently with the appearance of the apical membrane bilayer during cercaria to schistosomule transformation and remains on the surface throughout all life stages of the parasite in vertebrate hosts (Khayath et al. Reference Khayath, Vicogne, Ahier, BenYounes, Konrad, Trolet, Viscogliosi, Brehm and Dissous2007). Glucose uptake in schistosomes is mediated by PI3K/Akt/mTOR signal. The Akt protein, also known as protein kinase B, is associated not only with the expression of GTP4 but also with the shuttling of this transporter within the tegument (Abou-El-Naga Reference Abou-El-Naga2021; McKenzie et al. Reference McKenzie, Kirk and Walker2018; Morel et al. Reference Morel, Vanderstraete, Cailliau, Lescuyer, Lancelot and Dissous2014; Skelly and Shoemaker Reference Skelly and Shoemaker1996). Maharjan et al. (Reference Maharjan, Kirk, Lawton and Walker2023) demonstrated that lipid rafts could be crucial for glucose import into the parasite, potentially in response to host insulin.

Adult S. mansoni possesses two insulin receptors (IR1 and IR2) (Smp_341160; Smp_009990) (Khayath et al. Reference Khayath, Vicogne, Ahier, BenYounes, Konrad, Trolet, Viscogliosi, Brehm and Dissous2007). Both receptors are also present in schistosomula (Maharjan et al. Reference Maharjan, Kirk, Lawton and Walker2023). A potential insulin-like peptide has been identified in S. mansoni, although it is still unclear whether this peptide interacts with the IRs of S. mansoni (Wang et al. Reference Wang, Luo, Zhang, Yin, Dou and Cai2014). Each IR has a unique extracellular N-terminal domain that stabilizes the conformation around the bound ligand (Vicogne et al. Reference Vicogne, Pin, Lardans, Capron, Noël and Dissous2003). These receptors differ in essential signalling motifs and expression locations. IR1 is expressed at the tegumental basal membrane, in muscle tissues, and in the epithelial cells of the intestine, whereas IR2 is predominantly found in the parenchymal cells of adult worms (Khayath et al. Reference Khayath, Vicogne, Ahier, BenYounes, Konrad, Trolet, Viscogliosi, Brehm and Dissous2007). Invertebrates typically have a single IR that regulates both metabolism and growth, whereas vertebrates have two receptors: the IR and the insulin-like growth factor 1 receptor, which control glucose uptake and growth, respectively (Kim and Accili Reference Kim and Accili2002). Similarly, in Schistosoma, IR1 is likely specialized for glucose uptake, given its colocalization with glucose transporters GTP1 and GTP4 in the tegument. In contrast, the widespread expression of IR2 in the worm parenchyma suggests a role in growth control, similar to the insulin-like growth factor 1 receptor in mammals. Stimulation of IRs by human insulin activates the phosphatidylinositol 3 kinase/Akt/ mechanistic target of rapamycin (PI3K/Akt/mTOR) leading to increase the glucose uptake in S. mansoni (Abou-El-Naga Reference Abou-El-Naga2021; McKenzie et al. Reference McKenzie, Kirk and Walker2018). Several studies have highlighted the significance of schistosome insulin receptors for nutrition, growth, and reproduction (Abou-El-Naga Reference Abou-El-Naga2021; Elhenawy et al. Reference Elhenawy, Ashour, Nabih, Shalaby, El-Karef and Abou-El-Wafa2017; Vanderstraete et al. Reference Vanderstraete, Gouignard, Ahier, Morel, Vicogne and Dissous2013). RNA interference (RNAi) of the IRs decreases uptake and affects schistosome development (You et al. Reference You, Gobert, Cai, Mou, Nawaratna, Fang, Villinger and McManus2015).

Fibroblast growth factor receptors (FGFRs)

Secreted human fibroblast growth factor (FGF) interacts with surface RTKs of the FGF receptors (FGFRs) family, which includes four isoforms (FGFR1–FGFR4). FGF binds FGF ligand to FGFR, thereby phosphorylating mitogen-activated protein kinase (MAPK), phosphoinositide 3-kinase (PI3K)/AKT pathways. Human FGF1 (acidic FGF) and FGF2 (basic FGF) are the most active members of the FGF family and are universally expressed in human tissues. They are released in a signal peptide-independent manner (Ornitz and Itoh Reference Ornitz and Itoh2015).

In the genome of S. mansoni, two genes encoding FGFRs were identified. The fgfrA gene (Smp_175590) encodes a predicted protein with two extracellular immunoglobulin domains and a split tyrosine kinase domain, while the fgfrB gene (Smp_157300) product contains only one extracellular immunoglobulin domain. FGFRA and FGFRB are enzymatically active, expressed in the gonads of schistosomes, and upregulated following pairing, suggesting a role in parasite fertility (Hahnel et al. Reference Hahnel, Quack, Parker-Manuel, Lu, Vanderstraete, Morel, Dissous, Cailliau and Grevelding2014). The expression of fgfrA and fgfrB has been demonstrated in neoblast-like somatic stem cells, with evidence indicating that both receptors play a crucial role in maintaining schistosome stem cells (Wang et al. Reference Wang, Collins and Newmark2013). FGFRA is abundantly expressed in germinal/stem cells across various S. mansoni developmental stages including eggs, miracidia, cercariae, schistosomula, and adult worms. The distribution of FGFRA in embryonic cells of immature eggs and in the neural mass of mature eggs and miracidia, and its co-localization with stem cells in adult S. mansoni, strongly suggest its crucial roles in the maintenance of schistosome stem cells, in the development of the nervous and reproductive systems, and in the host-parasite interaction (Wang et al. Reference Wang, Collins and Newmark2013; Wendt and Collins Reference Wendt and Collins2016). In vitro, FGFRA of adult worms binds to human FGFs and activates the mitogen activated protein kinase (MAPK) pathway (Du et al. Reference Du, McManus, Fogarty, Jones and You2022). Inhibition of FGF signalling by the TK inhibitor significantly reduced egg hatching ability and altered the behavior of hatched miracidia from treated eggs, highlighting the critical role of FGF signalling in the life cycle of S. mansoni (Du et al. Reference Du, McManus, Fogarty, Jones and You2022). Inhibition of fgfrA in S. mansoni reduces stem cell signalling and increases cell apoptosis. Intravenous injection of mice with fgfrA-repressed eggs resulted in significantly smaller granulomas and a reduction in serum IgE levels, underscoring the crucial role of FGFRA in regulating the host immune response during schistosome infection (Du et al. Reference Du, McManus, French, Collinson, Sivakumaran, MacGregor, Fogarty, Jones and You2023).

Venus kinase receptors (VKRs)

S. mansoni expresses an unconventional family of receptor tyrosine kinases (RTKs) known as the Venus kinase receptor (VKR) family, which is unique to invertebrates and was first identified in S. mansoni. Typically, invertebrate genomes contain a single VKR gene, but platyhelminths possess two distinct copies of VKR (Vanderstraete et al. Reference Vanderstraete, Gouignard, Ahier, Morel, Vicogne and Dissous2013). VKRs have an extracellular Venus flytrap module, similar to the ligand-binding domain of class C G-protein coupled receptors, connected via a transmembrane segment to an intracellular tyrosine kinase (TK) domain. This Venus flytrap module has two lobes that close upon ligand binding (Vicogne et al. Reference Vicogne, Pin, Lardans, Capron, Noël and Dissous2003). In S. mansoni, the two receptors, VKR1 (Smp_019790) and VKR2 (Smp_153500), are activated by L-arginine and calcium ions, respectively (Gouignard et al. Reference Gouignard, Vanderstraete, Cailliau, Lescuyer, Browaeys and Dissous2012).

S. mansoni beta-integrin receptor Smβ-Int1 interacts with the SmVKRI. The three putative bridging molecule – SmILK, SmPINCH, and SmNck2 – mediate the Smβ-Int1/SmVKR1 cooperation. This process indicates that SmVKR1 can be activated in a ligand independent manner mediated by receptor/complex interaction (Gelmedin et al. Reference Gelmedin, Morel, Hahnel, Cailliau, Dissous and Grevelding2017).

VKRs are abundantly present in the germinal cells of miracidia, in the larval stage of the parasite, and in the oocytes within the ovary and oviduct of adult female worms. They play crucial roles in growth, differentiation, and reproduction through the PI3K/Akt/mTOR and mitogen-activated protein kinase (MAPK) pathways. VKR1 can also activate the c-Jun N-terminal kinase signal transduction pathway (Gouignard et al. Reference Gouignard, Vanderstraete, Cailliau, Lescuyer, Browaeys and Dissous2012; Vanderstraete et al. Reference Vanderstraete, Gouignard, Cailliau, Morel, Hahnel, Leutner, Beckmann, Grevelding and Dissous2014; Vicogne et al. Reference Vicogne, Pin, Lardans, Capron, Noël and Dissous2003). The Smβ-Int1/SmVKR1 signalling complex plays a crucial role in oocytes differentiation and survival of paired schistosomes (Gelmedin et al. Reference Gelmedin, Morel, Hahnel, Cailliau, Dissous and Grevelding2017).

Both S. mansoni VKR1 (Smp_019790) and VKR2 (Smp_153500) genes are highly transcribed in the ovaries of females compared to the testes of male worms, and each gene exhibits a distinct expression profile. The distribution of each VKR in S. mansoni correlates with its role in oocyte maturation. VKR1 is expressed in mature oocytes located in the posterior part of the ovary and is involved in oocyte migration. VKR2, on the other hand, is expressed in immature oocytes in the anterior part of the ovary and is responsible for their proliferation and growth (Gouignard et al. Reference Gouignard, Vanderstraete, Cailliau, Lescuyer, Browaeys and Dissous2012; Vanderstraete et al. Reference Vanderstraete, Gouignard, Cailliau, Morel, Hahnel, Leutner, Beckmann, Grevelding and Dissous2014). Mathavan et al. (Reference Mathavan, Liu, Robinson, El-Sakkary, Elatico, Gomez, Nellas, Owens, Zuercher, Navratilova, Caffrey and Beis2022) identified GSK1520489, GSK986310, GW696155, and SB- 710363 as kinase inhibitors of S. mansoni VKR2. They inhibit the enzymatic activity and induce phenotypic changes in the worm.

Transforming growth factor receptors

The TGF-β superfamily consists of a wide variety of structurally related polypeptide growth factors that are known to mediate numerous physiological processes, including growth and differentiation, cell death, and tissue repair (Chen et al. Reference Chen, Qin and Simons2023). Members of the TGF-β superfamily are divided into two major subfamilies based on sequence homology and the distinct downstream pathways they activate. The two subfamilies are the TGF-β/activin/nodal subfamily and the bone morphogenetic protein/growth and differentiation factor/Muellerian inhibiting substance (BMP/GDF/MIS) subfamily (Baba et al. Reference Baba, Rah, Bhat, Mushtaq, Parveen, Hassan, Hameed Zargar and Afroze2022).

When activated by a ligand, TGF-β family members bind to a group of transmembrane receptor serine/threonine kinases and transmit signals through them. The receptors are divided into two subtypes: transforming growth factor–β type I receptor (TβRI) (Smp_049760) and type II (TβRII) (Smp_144390). Type II receptor is crucial for ligand binding, and upon ligand binding, it activates type I receptor through phosphorylation (Osman et al. Reference Osman, Niles, Verjovski-Almeida and LoVerde2006). Activated TβRI subsequently transmits the signal to a member of the cytoplasmic Smad family, which can transport the signal to the nucleus and regulate the transcription of specific genes in response to the ligand (Freitas et al. Reference Freitas, Jung and Pearce2009). Smads are a group of proteins that act as intracellular signalling transducers for the TGF-β family. In mammals, the two subfamilies of the TGF-β superfamily activate different classes of Smad proteins. Members of the TGF-β subfamily activate Smad2 and Smad3 homologues, while members of the BMP subfamily activate Smad1, Smad5, and Smad8 homologues (Hata and Chen Reference Hata and Chen2016).

Several components of TGF-β signalling have been identified in S. mansoni, including the TGF-β transmembrane receptor serine/threonine kinases, also known as SmTβRI (Davies et al. Reference Davies, Shoemaker and Pearce1998) and SmTβRII (Osman et al. Reference Osman, Niles, Verjovski-Almeida and LoVerde2006), and two ligands of the schistosome TGF-β family: inhibin/Activin (SmInAct) (Freitas et al. Reference Freitas, Jung and Pearce2007) and SmBMP. SmBMP is expressed in the egg, cercariae, and the protonephridia of adult S. mansoni worm (Freitas et al. Reference Freitas, Jung and Pearce2009). The TGF-β signal also includes four Smad proteins (SmSmad1, SmSmad2, SmSmad4, SmSmad1B) (Carlo et al. Reference Carlo, Osman, Niles, Wu, Fantappie, Oliveira and LoVerde2007; Osman et al. Reference Osman, Niles and LoVerde2004), in addition to six scaffolding/regulatory proteins that play an important role in signal regulation in the TGF-β pathway. They comprise SmSARA (Verjovski-Almeida et al. Reference Verjovski-Almeida, DeMarco, Martins, Guimarães, Ojopi, Paquola, Piazza, Nishiyama, Kitajima, Adamson, Ashton, Bonaldo, Coulson, Dillon, Farias, Gregorio, Ho, Leite, Malaquias, Marques, Miyasato, Nascimento, Ohlweiler, Reis, Ribeiro, Sá, Stukart, Soares, Gargioni, Kawano, Rodrigues, Madeira, Wilson, Menck, Setubal, Leite and Dias-Neto2003), SmGCN5, SmCBP (Carlo et al. Reference Carlo, Osman, Niles, Wu, Fantappie, Oliveira and LoVerde2007), SmFKBP12 (Knobloch et al. Reference Knobloch, Rossi, Osman, LoVerde, Klinkert and Grevelding2004), SmeIF2α (McGonigle et al. Reference McGonigle, Beall and Pearce2002), and Sm14-3-3ε (McGonigle et al. Reference McGonigle, Beall, Feeney and Pearce2001).

SmTβRII and SmTβRI are expressed on the surface of the parasite, and in the case of SmTβRI, its expression is upregulated following infection of the mammalian host (Davies et al. Reference Davies, Shoemaker and Pearce1998). TβRII is also localized in the vitelline and gut epithelial cells of female worms and the sub-tegumental cells of male worms (Osman et al. Reference Osman, Niles, Verjovski-Almeida and LoVerde2006). Due to this localization, the schistosome TGF-β signalling pathway may play a crucial role in the development of vitelline cells in female worms and egg embryogenesis (LoVerde et al. Reference LoVerde, Osman and Hinck2007). SmInAct expression is closely associated with the reproductive potential of the parasite. RNAi-mediated knockdown of SmInAct in eggs halted their development, indicating that SmInAct plays a crucial role in embryogenesis (Freitas et al. Reference Freitas, Jung and Pearce2007; Osman et al. Reference Osman, Niles, Verjovski-Almeida and LoVerde2006). SmSmad4 is localized within the epithelia surrounding the gut and vitellarium, as well as in the sub-tegument and muscles of males (Osman et al. Reference Osman, Niles and LoVerde2004). SmSmad2 is found within the vitellarium, developing egg and ovary of the female worm, as well as in the testes and tubercles of the male worm (Osman et al. Reference Osman, Niles and LoVerde2001). Therefore, the TGF-β signalling pathway in S. mansoni has been implicated in host-parasite interactions, parasite reproductive development, and embryogenesis.

SmTβRII is able to activate SmTβRI in the presence of human TGF-β1, which subsequently activates SmSmad2 and promotes its interaction with SmSmad4, thereby facilitating the transfer of the signal from the receptor complex to the Smad proteins. The newly formed Smad complex translocates into the nucleus, where it associates with nuclear proteins that guide the complex to specific promoter sequences, regulating the transcription of target genes (Freitas et al. Reference Freitas, Jung and Pearce2009). Oliveira et al. (Reference Oliveira, Carvalho, Verjovski-Almeida and LoVerde2012) demonstrated that in vitro treatment with human TGF-β1 led to changes in expression levels of 381 S. mansoni genes, including 316 downregulated genes and 65 upregulated genes. Among these genes, there are genes related to morphology, development, and cell cycle that could influence effects of cytokine on the worm. Osman et al. (Reference Osman, Niles, Verjovski-Almeida and LoVerde2006) demonstrated that TGF-β signalling regulates the expression of the gynecophoral canal protein (SmGCP). This protein is located on the surface of the gynecophoric canal of the male where the female resides for sexual maturation. It is also found on the entire surface of females in copula, but not on unmated males or immature females (Bostic and Strand Reference Bostic and Strand1996). Therefore, it could be suggested that SmGCP might be a gene product induced by the TGF-β pathway and could serve as a crucial signalling molecule for worm pairing (Osman et al. Reference Osman, Niles, Verjovski-Almeida and LoVerde2006).

A distinctive biological trait of schistosomes is that sexual maturation of the female depends on continuous pairing contact with the male. After pairing, mitosis and differentiation are triggered in the female, leading to the development of reproductive organs, including the ovary and vitellarium, followed by the production of eggs (Kunz Reference Kunz2001). Eggs are important for propagation of the parasite life cycle and provoking pathogenesis. In S. mansoni, the TGF-β pathway is involved in female reproductive development and egg embryogenesis (Freitas el al. 2007; Knobloch et al. Reference Knobloch, Beckmann, Burmeister, Quack and Grevelding2007; Osman et al. Reference Osman, Niles, Verjovski-Almeida and LoVerde2006). As TβRs are exposed on the surface, this creates the potential for communication between the male and female parasites. In addition to utilizing host growth factors, it is suggested that schistosomes might also encode endogenous growth factor peptides that have a high degree of sequence similarity with their mammalian orthologues as developmental signals (LoVerde et al. Reference LoVerde, Osman and Hinck2007).

Nuclear hormone receptors (Table 2 )

Nuclear receptors (NRs) are crucial transcriptional regulators that control the expression of specific genes involved in animal development, differentiation, and reproduction. Regulation is achieved by controlling the transcription of target genes through binding to specific DNA response elements (Kunz Reference Kunz2001). NRs are part of a large protein superfamily that includes intracellular receptors for hydrophobic signalling molecules such as steroid hormones, thyroid hormones, and proteins activated by intracellular metabolites (Wu and LoVerde Reference Wu and LoVerde2019). During the development of schistosomes in their hosts, several hormonal signals may be derived from the schistosome itself or from the host and exert this control through nuclear receptors. As mentioned previously, sexual maturation of female schistosomes depends on continuous pairing contact with the male leading to the development of the reproductive organs and production of eggs (Kunz Reference Kunz2001). Male worms also react to the excretory-secretory products of female worms (Childs et al. Reference Childs, Shirazian, Gloer and Schiller1986). Although mammalian sex hormones have no direct effect on the fertility of paired adult schistosome worm maintained in culture (Morrison et al. Reference Morrison, Vande Waa and Bennett1986), however, estrogens and androgens influence worm survival in the host (Escobedo et al. Reference Escobedo, Roberts, Carrero and Morales-Montor2005).

Table 2. Nuclear hormone receptors and nuclear transport receptors

Gene ID is extracted from the WormBase ParaSite using the reference genome for S. mansoni, SM_V10 (WormBase ParaSite 2024).

Typically, NRs share a common structure consisting of A/B, C, D, E, F domains and N and C terminals. The N-terminal A/B domain is highly variable and is regulated by interaction with coregulatory proteins and also contains a ligand independent activation function (AF-1), while the C domain, known as the DNA-binding domain (DBD), is the most conserved region, featuring two zinc finger motifs, and is responsible for NR binding to specific DNA sequences. They provide sequence-specific DNA recognition to the regulatory region of the target gene called the hormone response element (HRE). The conserved sequence of the first zinc finger contains a motif called P-box, which is responsible for binding to the target gene, while the conserved sequence of the second zinc finger with a motif called D-box is involved in dimerization. The D domain is poorly conserved and acts as a flexible hinge between DBD and ligand-binding (LBD) domains, giving them some independent mobility. The E domain contains the LBD that controls receptor activity by binding to other LBDs and interacting directly with co-regulatory proteins. It also contains the dimerization surface and a ligand-dependent transcription activation domain (AF-2) (Simons et al. Reference Simons, Edwards and Kumar2014). The F domain is sometimes included as part of domain E (E/F domain) (Patel and Skafar Reference Patel and Skafar2015; Schote Reference Schote2007) (Figure 1a).

Figure 1. (a) Schematic diagram of a typical nuclear receptor (NR) consisting of A/B, C, D, E, F domains and N and C terminals. The A/B domain includes the activation function 1 (AF-1), C domain is a DNA binding domain (DBD), D domain is a hinge region, E domain contains a ligand binding domain (LBD), and AF-2. (b) Schematic diagram for atypical NRs of S. mansoni containing two DBDs and a single LBD.

Atypical NRs are found in certain animals. In arthropods and nematodes, some NRs have a DBD but lack an LBD. In contrast, vertebrates have NRs that lack a DBD but contain an LBD. In addition, the most notable outcome of identifying NRs in S. mansoni leads to the discovery of three new members, each of which contains a distinctive combination of two DBDs arranged in tandem with a single LBD (Wu et al. Reference Wu, Niles, El-Sayed, Berriman and LoVerde2006; Wu et al. Reference Wu, Niles, Hirai and LoVerde2007a). This is followed by further identification of this member in other invertebrates (Wu et al. Reference Wu, Niles and LoVerde2007b; Wu and LoVerde Reference Wu and LoVerde2023).

Based on phylogenetic reconstructions of the DBD and LBD, NRs are divided into six classical subfamilies (NR1-NR6). In addition, an extra subfamily, NR0, has been identified. Members of this subfamily either contain only a DBD (NR0A) or only an LBD (NR0B) (Nuclear Receptors Nomenclature Committee 1999). S. mansoni contains 21 NRs that can be categorized into many subfamilies, including NR1, NR2, NR4, and NR5 (Wu and LoVerde Reference Wu and LoVerde2019). NRs in S. mansoni consist of six members in subfamily 1 (NR1), nine members in subfamily 2 (NR2), one member in subfamily 4 (NR4), and two members in subfamily 5 (NR5). S. mansoni contains also three novel members, each characterized by a distinctive combination of two DBDs in tandem with LBD – Sm2DBD-NRα, Sm2DBD-NRβ, and Sm2DBD-NRγ with a novel modular structure: A/B-DBD-DBD-hinge-LBD organization (Wu et al. Reference Wu, Niles, El-Sayed, Berriman and LoVerde2006; Wu et al. Reference Wu, Niles, Hirai and LoVerde2007a). The worm does not contain NRs in subfamily 3 (NR3) or subfamily 6 (NR6). Among these 21 receptors, the full-length cDNA of 14 members has been isolated and studied (Wu et al. Reference Wu, Niles, El-Sayed, Berriman and LoVerde2006).

The control function of the NRs on gene expression often requires interaction with endogenous or exogenous ligands. Unlike other transcription factors, NRs can modulate their activity by binding to specific ligands, which are primarily small lipophilic molecules that readily penetrate biological membranes (Novac and Heinzel Reference Novac and Heinzel2004). This binding creates a direct link between cellular signals and the transcriptional responses of the cell. These lipophilic ligands include fatty acids, steroids, retinoids, phospholipids, vitamin D, and thyroid hormone. However, NRs without known ligands have yet been identified and referred to as orphan receptors (Li et al. Reference Li, Weth, Haimann, Möscheid, Huber and Grevelding2024). NRs execute their gene regulatory role by binding to the regulatory regions of target genes (often called hormone response elements (HREs) to activate or repress mRNA synthesis. Binding occurs following ligand-induced activation and subsequent recruitment of co-factors. NRs can bind to HREs as homodimers, heterodimers, or monomers (Weikum et al. Reference Weikum, Liu and Ortlund2018). Response elements are composed of distinct arrangements of the core motif that can be recognized by NRs derivatives of this same DNA core motif (Pawlak et al. Reference Pawlak, Lefebvre and Staels2012). HRE-like elements have been identified in several schistosome genes, and gel-shift assays have shown that nuclear proteins can bind to these sequences. Notably, the F10 gene, encoding the F10 egg shell protein, contains a monomeric HRE-like element, and the protein binding pattern to the F10 promoter is modified by the estrogen antagonist tamoxifen (Giannini et al. Reference Giannini, Caride, Braga and Rumjanek1995).

NRs regulate transcription by binding to the promoter region of their target gene via the DBD and regulating the expression of related target genes through the recruitment of coactivators or corepressors when ligands bind to the receptors (Hong et al. Reference Hong, Pan, Guo, Xu and Zhai2019). NRs can be classified into two broad subtypes based on their mechanisms of action. Type I NRs are found in the cytoplasm in the absence of ligands, where they form complexes with heat shock proteins (HSPs) that regulate their cellular localization, protein stability, and transcriptional activity (Echeverria and Picard Reference Echeverria and Picard2010). When a ligand binds, the receptor is released from the HSP, undergoes dimerization, and trans-locates to the nucleus. In the nucleus, the ligand-receptor complex associates with coactivators and RNA polymerase, enabling binding to and activation of target genes (Bulynko and O’Malley Reference Bulynko and BW2011). Type II NRs are located in the nucleus, bound to DNA, regardless of their ligand-binding status. These receptors typically form heterodimers with retinoid X receptors (RXRs). In the absence of a ligand, the NR is associated with corepressor proteins. Ligand binding to the NR triggers the dissociation of corepressors and the recruitment of coactivator proteins, which then attract RNA polymerase. This complex facilitates the transcription of downstream DNA into RNA, ultimately leading to protein production and changes in cellular function (Lazar Reference Lazar2017).

S. mansoni nuclear receptors in subfamily 1

The well-characterized proteins of S. mansoni NR1 are S. mansoni thyroid receptors (SmTRs). Two homologs of vertebrate TR have been identified in S. mansoni (SmTRα and SmTRβ) (Smp_134490, Smp_174260). Phylogenetic analysis indicates that these two copies resulted from a gene duplication specific to Schistosoma (Wu et al. Reference Wu, Niles, Hirai and LoVerde2007a). Both proteins exhibit the consensus structure of TR, featuring a conserved N-terminal signature in the A/B domain typical of TRs, along with the specific CEGCKGFFRR sequence of the NR1 subfamily. Like vertebrate members of this family, SmTRs can form a dimer with retinoid X receptor (SmRXR1). SmTRs could bind to vertebrate TR core DNA elements as a monomer, a homodimer, or a heterodimer by binding with RXR (Wu et al. Reference Wu, Niles, El-Sayed, Berriman and LoVerde2006; Wu et al. Reference Wu, Niles, Hirai and LoVerde2007a). Thyroid hormone (TH) binds to the LBD of the TR, inducing a conformational change in C-terminus of the receptor. This change causes the dissociation of corepressors from the TR, allowing coactivators to bind to the C-terminus in a hormone-dependent manner. The TR and coactivator complex then activates target gene expression (Lazar Reference Lazar2017).

An ortholog of the Drosophila ecdysone-induced protein 78 has been identified in S. mansoni (Wu et al. Reference Wu, Niles, El-Sayed, Berriman and LoVerde2006; Wu et al. Reference Wu, Tak and LoVerde2008). It is directly involved in ecdysone signaling. SmE78 (Smp_000340) is expressed throughout schistosome development, with the highest expression levels observed in the miracidia and egg stages (Wu et al. Reference Wu, Tak and LoVerde2008). Nirde et al. (Reference Nirde, Torpier, De Reggi and Capron1983) have shown that S. mansoni can synthesize the steroid hormone ecdysone. Ecdysterone effectively stimulates host location activities in miracidia (Shiff and Dossaji Reference Shiff and Dossaji1991). However, it remains to be demonstrated whether ecdysone-induced protein 78 is involved in the transduction of an ecdysone signal in S. mansoni (Wu and LoVerde Reference Wu and LoVerde2011).

Another potential member of the SmNP1 subfamily is Smp_248100, an uncharacterized protein from S. mansoni. Primary sequence analysis has confirmed that Smp_248100 contains a DNA-binding domain (DBD) with high sequence similarity to DBDs of other vertebrate and invertebrate NRs, including HR96 from Drosophila melanogaster and DAF-12 from Caenorhabditis elegans. SmHR96α and SmHR96β are homologues of Drosophila hormone receptor 96 (DHR96) (Hu et al. Reference Hu, Wu, Niles and LoVerde2006a; Wu et al. Reference Wu, Niles, El-Sayed, Berriman and LoVerde2006). SmHR96α interacts with SmRXR1 (Hu et al. Reference Hu, Niles and LoVerde2006b). The mRNA of SmHR96α is expressed throughout every stage of the S. mansoni life cycle, with particularly high expression levels in eggs and cercariae. The SmHR96α protein is located in subtegumental and parenchymal cells in both male and female worms, as well as in the ovaries, eggs, and vitelline cells of mature female worms (Hu et al. Reference Hu, Wu, Niles and LoVerde2006a). SmHR96β is recently named Vitellogenic Factor 1. This factor is important in vitelline cell development, and this NR is important for female sexual development after pairing with a male worm (Wang et al. Reference Wang, Chen and Collins2019).

The divergent member SmNR1 is a member of NR subfamily I with no known orthologue. The gene of this member is located on chromosome 1 of S. mansoni and is highly expressed in eggs, sporocysts, and juvenile worms (Wu et al. Reference Wu, Niles, Hirai and LoVerde2007c). The divergent member SmNR1 is a partner of SmRXR1. It requires RXR to form a heterodimer that confers binding to hormone response element (Kojetin et al. Reference Kojetin, Matta-Camacho, Hughes, Srinivasan, Nwachukwu, Cavett, Nowak, Chalmers, Marciano, Kamenecka, Shulman, Rance, Griffin, Bruning and Nettles2015). Mutagenesis analysis indicates that SmCBP1, a co-regulatory protein known to interact with SmFTZ-F1 (Bertin et al. Reference Bertin, Oger, Cornette, Caby, Noël, Capron, Fantappie, Rumjanek and Pierce2006), can mediate interactions with the LBD of SmRXR1 and SmNR1 (Fantappié et al. Reference Fantappié, Bastos de Oliveira, de Moraes Maciel, Rumjanek, Wu and LoVerde2008a). The upstream region of the p14 gene has a novel NR response element containing DNA core motif, composed of an atypically spaced direct repeat 17. SmRXR1 and SmNR1 divergent members specifically bound to the p14-direct repeat 17 element as a heterodimer. SmRXR1, but not SmNR1, is bound to the motif as a monomer. The expression of the S. mansoni p14 gene, which is an eggshell precursor gene expressed only in the vitelline cells of sexually mature female worms in response to an as yet unidentified male stimulus, is regulated through NR signalling pathway (Fantappié et al. Reference Fantappié, Furtado, Rumjanek and LoVerde2008b).

S. mansoni nuclear receptors in subfamily 2

S. mansoni NRs in subfamily 2 include SmTR2/4, 9-cis-retinoic acid receptors (RXR), and hepatocyte nuclear factor 4 (HNF4). The SmTR2/4 gene is expressed in all developmental stages of S. mansoni with a higher level in cercaria and may play a role in regulating female reproductive development (Hu et al. Reference Hu, Wu, Niles and LoVerde2006c).

SmRXR1 (Smp_097700) and SmRXR2 (Smp_073470) are the vertebrate RXR homologues present in S. mansoni (de Mendonça et al. Reference de Mendonça, Escriva, Bouton, Zelus, Vanacker, Bonnelye, Cornette, Pierce and Laudet2000; Freebern et al. Reference Freebern, Niles and LoVerde1999a; Freebern et al. Reference Freebern, Osman, Niles, Christen and LoVerde1999b). Both SmRXRs originated from a Schistosoma-specific gene duplication, like SmTRs. RXR is involved in multiple signalling pathways within the cell nucleus, and its heterodimers control the activity of other NRs (Bertin et al. Reference Bertin, Caby, Oger, Sasorith, Wurtz and Pierce2005). SmRXR1 can form heterodimers with SmTRα, SmTRβ (Wu et al. Reference Wu, Niles and LoVerde2007b), SmHR96α (Hu et al. Reference Hu, Niles and LoVerde2006b), and SmNR1 divergent member (Wu et al. Reference Wu, Niles, Hirai and LoVerde2007c). It can also bind to the cis-elements of the S. mansoni p14 gene (Fantappié et al. Reference Fantappié, Furtado, Rumjanek and LoVerde2008b; Freebern et al. Reference Freebern, Osman, Niles, Christen and LoVerde1999b), which is regulated by a stimulus from the male schistosome (LoVerde et al. Reference LoVerde, Niles, Osman and Wu2004). SmRXR1 mRNA is consistently expressed throughout the developmental stages (Fantappié et al. Reference Fantappié, Furtado, Rumjanek and LoVerde2008b). The co-regulatory protein SmCBP1 can mediate interactions with both SmRXR1 and SmNR1 (Fantappié et al. Reference Fantappié, Bastos de Oliveira, de Moraes Maciel, Rumjanek, Wu and LoVerde2008a). In contrast, SmRXR2 fails to form a heterodimer with SmTRα or SmTRβ (Wu et al. Reference Wu, Niles and LoVerde2007b), SmHR96α (Hu et al. Reference Hu, Niles and LoVerde2006b) or SmNR1 divergent member (Wu et al. Reference Wu, Niles, Hirai and LoVerde2007c). SmRXR2 mRNA is expressed at all life cycle stages, with higher levels in cercariae and miracidia – the free-living larval stages (de Mendonça et al. Reference de Mendonça, Escriva, Bouton, Zelus, Vanacker, Bonnelye, Cornette, Pierce and Laudet2000; Freebern et al. Reference Freebern, Niles and LoVerde1999a). However, the protein expression differs significantly from mRNA, showing high levels in schistosomula but much lower levels in cercariae and miracidia (de Mendonça et al. Reference de Mendonça, Escriva, Bouton, Zelus, Vanacker, Bonnelye, Cornette, Pierce and Laudet2000).

HNF4 is a class of NRs responsible for regulation of gluconeogenesis, bile acid synthesis, cholesterol, and lipid metabolism in the liver of mammals (Chen et al. Reference Chen, Vasoya, Toke, Parthasarathy, Luo, Chiles, Flores, Gao, Bonder, Su and Verzi2020). Stem cells from the blood-digesting gut of S. mansoni express the hnf4 gene (Smp_174700), identified by single-cell sequencing. RNAi assay revealed the importance of the hnf4 gene for gut maintenance, nutrient digestion, and pathology induction, and indirectly showed its importance for parasite growth (Wendt et al. Reference Wendt, Zhao, Chen, Liu, O’Donoghue, Caffrey, Reese and Collins2020). In S. japonicum, HNF4 expression is higher in female than in male worms, both at transcriptional and protein levels. HNF4 is expressed in the reproductive system and intestinal tissues of worms, as well as in cercariae and eggs (Wu et al. Reference Wu, Huang, Zhao, Umar, Chen, Yu and Huang2024). Furthermore, HNF4 plays an important role in blood feeding and interaction with vital pathways such as glucose, lipid, and nucleotide metabolism. Schistosomes obtain hemoglobin, plasma proteins, and immunoglobulins from the blood to meet their energy needs. The processing of these proteins is carried out by a complex system comprising various proteases, many of which are associated with HNF4 in S. japonicum. Furthermore, HNF4 is connected to several proteins involved in carbohydrate metabolism (Wu et al. Reference Wu, Huang, Zhao, Umar, Chen, Yu and Huang2024). Glucose, an essential nutrient, serves as the main energy source for schistosomes, providing them with the energy necessary for their growth and reproduction (You et al. Reference You, Stephenson, Gobert and McManus2014). Adult worms possess a higher lipid content and rely on their host for lipid acquisition (Skelly et al. Reference Skelly, Da’dara, Li, Castro-Borges and Wilson2014). A strong correlation between SjHNF4 and the phospholipid metabolism pathway suggests that SjHNF4 also contributes to lipid metabolism in S. japonicum (Wu et al. Reference Wu, Huang, Zhao, Umar, Chen, Yu and Huang2024).

S. mansoni nuclear receptors in subfamily 4

NR4A is the only member of subfamily 4 identified in Schistosoma worms. SmNR4A, like the human and Drosophila members of NR subfamily 4, has an atypical LBD suggesting that SmNR4A is an orthologue of Drosophila and human NR4A (Wu and LoVerde Reference Wu and LoVerde2021). SmNR4A is highly expressed in daughter sporocysts and adult worms, but scarcely in cercariae and early schistosomules (Wu and LoVerde Reference Wu, Tak and LoVerde2008).

S. mansoni nuclear receptors in subfamily 5

Fushi tarazu-factor 1 (FTZ-F1) (Smp_328000) NRs are the only receptors of subfamily 5 that have been characterized in S. mansoni. SmFTZ-F1 NRs contain two NRs: the SmFTZ-F1α belonging to NR5A group and the SmFTZ-F1 belonging to NR5B group (Lu et al. Reference Lu, Wu, Niles and LoVerde2006; Wu et al. Reference Wu, Niles, El-Sayed, Berriman and LoVerde2006). Smftz-f1α is continuously expressed throughout the schistosome life cycle, with the highest expression level observed at the egg stage (Lu et al. Reference Lu, Wu, Niles and LoVerde2006). RT-PCR reveals that Smftz-f1 is expressed at all developmental stages, with higher mRNA levels in miracidia, sporocysts, and cercariae. However, protein expression levels differ, being highest in cercariae, schistosomula, and male worms (de Mendonça et al. Reference De Mendonça, Bouton, Bertin, Escriva, Noël, Vanacker, Cornette, Laudet and Pierce2002). Romero et al. (Reference Romero, Cobb, Collins, Kliewer, Mangelsdorf and Collins2021) identified the micro-exon gene meg-8.3 as a target gene of SmFtz-F1, and this gene is expressed exclusively in the esophageal gland of the worm. They also found that Smftz-f1 and meg-8.3 are essential for maintaining the esophageal gland and preserving the integrity of the worm’s head. The esophageal gland plays a crucial role in protecting the worm from host attacks (Lee et al. Reference Lee, Chong and Newmark2020).

S. mansoni nuclear receptors subfamily with two DBDs and a single LBD (2DBD-NRs) (Figure 1b)

One of the most interesting findings is the isolation of a new group of NRs from S. mansoni, in which each receptor contains two DBDs and a single LBD (2DBD-NRs) (Wu et al. Reference Wu, Niles, El-Sayed, Berriman and LoVerde2006). S. mansoni Genome Project verified its presence (Berriman et al. Reference Berriman, Haas, LoVerde, Wilson, Dillon, Cerqueira, Mashiyama, Al-Lazikani, Andrade, Ashton, Aslett, Bartholomeu, Blandin, Caffrey, Coghlan, Coulson, Day, Delcher, DeMarco, Djikeng, Eyre, Gamble, Ghedin, Gu, Hertz-Fowler, Hirai, Hirai, Houston, Ivens, Johnston, Lacerda, Macedo, McVeigh, Ning, Oliveira, Overington, Parkhill, Pertea, Pierce, Protasio, Quail, Rajandream, Rogers, Sajid, Salzberg, Stanke, Tivey, White, Williams, Wortman, Wu, Zamanian, Zerlotini, Fraser-Liggett, Barrell and El-Sayed2009). These NRs have a novel modular structure: A/B-DBD-DBD-hinge-LBD organization in the NR. Sm2DBD-NRα is able to form a homodimer but cannot form a heterodimer with RXRs. S. mansoni expresses three 2DBD-NRs (Sm2DBD-NRα, Sm2DBD-NRβ, and Sm2DBD-NRγ) located on different chromosomes (Wu et al. Reference Wu, Niles, Hirai and LoVerde2007a). 2DBD-NRs have been identified and/or isolated only in Platyhelminths (Wu et al. Reference Wu, Niles, El-Sayed, Berriman and LoVerde2006; Wu et al. Reference Wu, Niles, Hirai and LoVerde2007a; Wu and LoVerde 2021) and Mollusca (Kaur et al. Reference Kaur, Jobling, Jones, Noble, Routledge and Lockyer2015; Vogeler et al. Reference Vogeler, Galloway, Lyons and Bean2014), suggesting they may be species-specific. Recently, 2DBD-NRs were identified in different animals (Wu and LoVerde Reference Wu and LoVerde2023). As shown by qRT-PCR, the three Sm2DBD NRs are developmentally regulated. Sm2DBD-NRα was found in sporocysts, cercariae, schistosomules, and male and female worms; Sm2DBDNRβ was expressed at high levels in eggs, sporocysts, cercariae, and male worms; and Sm2DBD-NRγ was only found in cercariae and juvenile worms (Wu et al. Reference Wu, Niles, Hirai and LoVerde2007a).

Nuclear transport receptors (Table 2)

Nuclear transport is the mechanism by which molecules move across the nuclear membrane of a cell. Transport of proteins and RNA across the nucleus occurs through the nuclear pore complex and is facilitated by a superfamily of transport receptors collectively known as karyopherins. The entry and exit of the molecules from the nucleus is tightly controlled by the nuclear pore complexes (NPCs). Although small molecules can enter the nucleus without regulation, macromolecules such as RNA and proteins require association with nuclear transport receptors (Mackmull et al. Reference Mackmull, Klaus, Heinze, Chokkalingam, Beyer, Russell, Ori and Beck2017).

Transport receptors that import cargo are called importins, and transport receptors that export cargo are called exportins. Exportins (XPOs) are nuclear export receptors concerned with export of various RNA species generated in the nucleus to the cytoplasm via the NPCs. This transport is vital for gene expression in eukaryotic cells. The nucleocytoplasmic transport occurs through different mechanisms: small RNAs (such as tRNAs and microRNAs) bind directly to export receptors, while larger RNAs (including ribosomal RNAs and mRNAs) use a more complex process. XPOs bind nuclear cargo only by identifying short signal peptides on cargo proteins or specific motifs on RNA cargoes (Köhler and Hurt Reference Köhler and Hurt2007). Furthermore, XPOs export only functional mRNAs into the cytoplasm. This quality control step is an important step, as faulty or unprocessed mRNAs can be harmful if translated in the cytoplasm (Lackner and Bähler Reference Lackner and Bähler2008).

Eight XPOs have been characterized (XPOs1-7), in addition to XPOT (Mingot et al. Reference Mingot, Bohnsack, Jäkle and Görlich2004). Homologs of XPO1 (Smp_124820), XPO5 (Smp_152800), and XPOT (Smp_137650) are present in animals, fungi, and plants, while nematodes and arthropods lose XPO5 or XPO1 during evolution (Murphy et al. Reference Murphy, Dancis and Brown2008). Schistosoma has a complex life cycle, and several life cycle stages are present in different hosts and environments, thus indicating differential gene regulation. Abreu et al. (Reference Abreu, Pereira, Oliveira, Gomes Mde, Jannotti-Passos, Borges and Guerra-Sá2013) identified the presence of XPO5, XPOT, and XPO1 at various stages of the S. mansoni life cycle, suggesting that exportins play a key role in the transport of different RNAs. Moreover, the authors demonstrated that XPOs are upregulated in schistosomula more than cercariae. As the level of protein synthesis is increased in schistosomula during the first 24 h after transformation (Blanton and Licate Reference Blanton and Licate1992), this may involve an alteration in protein synthesis (Abreu et al. Reference Abreu, Pereira, Oliveira, Gomes Mde, Jannotti-Passos, Borges and Guerra-Sá2013). XPO1 was found to be the most expressed receptor in all stages of life cycle of schistosomes compared to XPO5 and XPOT (Abreu et al. Reference Abreu, Pereira, Oliveira, Gomes Mde, Jannotti-Passos, Borges and Guerra-Sá2013). XPO5 is involved in the export of microRNAs, while XPOT is involved in the export of tRNAs. XPO1 plays a crucial role in the transport of several proteins with leucine-rich nuclear export signals: snRNAs involved in splicing, rRNA subunits, and certain mRNAs (Yang et al. Reference Yang, Guo, Chen, Gong, Jia and Sun2023). Thus, it is suggested that RNA transport by exportins may regulate cellular processes during the development of cercariae, schistosomula, and adult worms (Abreu et al. Reference Abreu, Pereira, Oliveira, Gomes Mde, Jannotti-Passos, Borges and Guerra-Sá2013).

Nuclear transport receptors are regulated by the small GTPase, Ran. Importins bind to the cargo protein that carries components of nuclear export signal (NES) into the cytoplasm through NPCs, and the cargo is released into the nucleus after transport, a process that is triggered by the binding of RanGTP (Tran et al. Reference Tran, King and Corbett2014).

XPOs form a complex with RanGTP to be translocated to the cytoplasm. In the cytoplasm, RanGDP dissociates the complex upon hydrolysis of RanGTP, resulting in the release of the cargo (Köhler and Hurt Reference Köhler and Hurt2007). XPO5 is responsible for exporting precursor miRNAs across the nuclear membrane into the cytoplasm and is therefore a critical step in miRNA biogenesis. The pre-miRNAs are transported from the nucleus to the cytoplasm, where they are enzymatically processed to become mature miRNAs. However, miRNAs can also use XPO1 for nuclear-cytoplasmic shuttling (Castanotto et al. Reference Castanotto, Lingeman, Riggs and Rossi2009). Both XPO5 and XPOT bind directly to pre-miRNA and tRNA, respectively, in a RanGTP dependent manner and diffuse into the cytoplasm through the NPC, where the complex dissociates (Köhler and Hurt Reference Köhler and Hurt2007; Okada et al. Reference Okada, Yamashita, Lee, Shibata, Katahira, Nakagawa, Yoneda and Tsukihara2009). Ran-GTP is hydrolyzed, forming a Ran-GDP complex, which is then transported back to the nucleus. Thus, while importins rely on RanGTP to release their cargo, exportins need RanGTP to bind theirs. XPO1 does not directly interact with the snRNA, rRNA, and mRNA cargo proteins, but requires the cap-binding complex (CBC) protein and a NES containing adaptor protein and RanGTP to be released into the cytoplasm (Köhler and Hurt Reference Köhler and Hurt2007). Transport of different S. mansoni RNAs by XPO5, XPOT, and XPO1 is illustrated in Figure 2.

Figure 2. Schematic diagram of S. mansoni RNAs transport by XPO5, XPOT, and XPO1. XPO5 (Smp_152800) binds to pre-miRNA, and XPOT (Smp_137650) binds to tRNA directly in a RanGTP dependent manner. Once RNA proteins are exported to the cytoplasm, the RanGTP is converted to GDP resulting in the release of pre-miRNA from the XPO5 and tRNA from the XPOT. Nuclear export of snRNA, rRNA, and mRNA by XPO1 (Smp_124820) requires the cap-binding complex (CBC) protein, a nuclear export signal (NES) containing adaptor protein and RanGTP. The complex transits to the cytoplasm through the nuclear pore complex (NPC) and releases the cargo protein upon the hydrolysis of RanGTP.

Neurotransmitter receptors (Table 3 )

The schistosome nervous system is fundamental to the successful migration of the parasite through the host, as well as its feeding and egg-laying activities. The central nervous system of trematodes includes two pairs of cerebral ganglia, each of which is a bi-lobed structure. From each lobe of the cerebral ganglia extend pairs of dorsal, ventral, and lateral nerve cords. These longitudinal nerve cords are interconnected by transverse commissures along the length of the worm. Trematodes also possess a peripheral nervous system consisting of finer nerve fibers and plexuses. These connect to all major body structures, including the somatic musculature, the tegument, the oral and ventral suckers, the reproductive organs, and the alimentary tract. In addition, the surface of the worm is abundant in sensory nerve endings that act as an interface between the parasite and the host environment (Halton and Maule Reference Halton and Maule2004).

Table 3. Neurotransmitter receptors

Gene ID is extracted from the WormBase ParaSite using the reference genome for S. mansoni, SM_V10 (WormBase ParaSite 2024).

The schistosome nervous system is involved in signal transduction through synaptic and paracrine mechanisms, since schistosomes lack a circulatory system and therefore cannot carry out classical endocrine signalling (El-Shabasy et al. Reference El-Shabasy, Saleh, Said and Reda2024; Halton and Maule Reference Halton and Maule2004). Neurotransmitters bind to their cognate receptors and elicit effects directly or through second messenger cascades (Ribeiro and Geary Reference Ribeiro and Geary2010; Ribeiro et al. Reference Ribeiro, Gupta and El-Sakkary2012). Neurotransmitter receptors can be categorized into two main classes: Cys-loop ligand-gated ion channels and metabotropic, seven-transmembrane G protein-coupled receptors (GPCRs). Neurotransmitters include acetylcholine (ACh), glutamate, and the biogenic amines. Biogenic amines are a group of structurally related amino acid derivatives that serve as neurotransmitters across various organisms. This group includes catecholamines synthesized from tyrosine (dopamine, noradrenaline, adrenaline), serotonin from tryptophan (5-hydroxytryptamine (5-HT)), and histamine form histidine. In addition, biogenic amines include octopamine and its precursor tyramine. Octopamine is a tyrosine-derived and invertebrate-specific neurotransmitter (El-Sakkary et al. Reference El-Sakkary, Chen, Arkin, Caffrey and Ribeiro2018). In flatworms, including S. mansoni, biogenic amines play a crucial role in regulating muscle contraction and movement, activities essential for survival of the parasite within the host (Cheng et al. Reference Cheng, Li, Qin, Xu, Zhang, Liu, Gu and Jin2019; El-Sakkary et al. Reference El-Sakkary, Chen, Arkin, Caffrey and Ribeiro2018; Ribeiro and Geary, Reference Ribeiro and Geary2010). The most studied of these amines is serotonin, which causes muscle excitation in all flatworm species examined so far. Serotonin is widely distributed in S. mansoni nervous system, with evidence of a serotonin transport system in the worm (Patocka and Ribeiro Reference Patocka and Ribeiro2007). In addition to serotonin, flatworms possess both dopamine and histamine in their nervous systems. Dopamine, in particular, plays significant neuromuscular roles, which can be either excitatory or inhibitory depending on the flatworm species. In S. mansoni, dopamine induces relaxation of the body wall muscles, possibly by activating a receptor associated with neuromuscular structures (Taman and Ribeiro Reference Taman and Ribeiro2009). In addition to their motor effects, biogenic amines have been implicated in the regulation of metabolic activity in several flatworms (Caveney et al. Reference Caveney, Cladman, Verellen and Donly2006). Moreover, serotonin and dopamine are involved in the transformation of S. mansoni miracidia to the sporocyst stage (Ribeiro and Patocka Reference Ribeiro and Patocka2013; Taft et al. Reference Taft, Norante and Yoshino2010;), suggesting a probable role in parasite development.

G protein-coupled receptors (GPCRs)

GPCRs represent the largest family of trans-membrane receptors involved in cellular communication in living organisms. These receptors can detect extracellular signalling molecules such as ions, light, hormones, neurotransmitters, amino acids, and neuropeptides, subsequently initiating a series of intracellular signal transduction pathways to produce the corresponding physiological effects (Weis and Kobilka Reference Weis and Kobilka2018). Biogenic amines exert their effects by interacting with cell-surface receptors, most of which belong to the superfamily of GPCRs. These receptors play roles in various biological processes, including growth, differentiation, neuronal signaling, olfaction, metabolism, and reproduction. The significance of GPCRs is underscored by their medical importance, as 30% to 50% of all pharmaceutical compounds target GPCRs and the signalling pathways they mediate (Miao and McCammon Reference Miao and McCammon2016).

GPCRs are composed of an extracellular N-terminus, a bundle of seven transmembrane α-helices (7TM), connected by extracellular and intracellular loops, and an intracellular C-terminus. The extracellular region, which includes the N-terminus, is responsible for ligand binding and varies in size from relatively short and often unstructured sequences in rhodopsin-like receptors to larger globular domains in other GPCR classes (Lagerström and Schiöth Reference Lagerström and Schiöth2008). The intracellular region interacts with G proteins, arrestins, and other downstream effectors (Tobin et al. Reference Tobin, Butcher and Kong2008).

GPCRs can activate guanine nucleotide-binding proteins (G proteins), which are responsible for signal transduction within the cell. G proteins transmit signals within the cell by interacting with various effector molecules, typically leading to changes in second messenger concentrations and subsequent cellular responses (Luttrell Reference Luttrell2008). In addition, GPCRs can activate G protein-independent signalling pathways via adaptor proteins such as arrestins (Bologna et al. Reference Bologna, Teoh, Bayoumi, Tang and Kim2017). Moreover, GPCRs can collaborate with other membrane proteins such as integrins and receptor tyrosine kinases (RTKs) (Cattaneo et al. Reference Cattaneo, Guerra, Parisi, De Marinis, Tafuri, Cinelli and Ammendola2014; Pyne and Pyne Reference Pyne and Pyne2011). When a ligand binds to a GPCR, it undergoes conformational changes and releases membrane-associated G-protein subunits (α, β, and γ). In this activated state, the GPCR functions as a guanine nucleotide exchange factor, promoting the exchange of GDP for GTP in the α subunit, leading to its dissociation from the βγ dimer. Both the dissociated α subunit and the βγ dimer can then trigger downstream signalling pathways (Frooninckx et al. Reference Frooninckx, Van Rompay, Temmerman, Van Sinay, Beets, Janssen, Husson and Schoofs2012).

According to the GRAFS classification, the mammalian GPCR are classified into five main families; Rhodopsin (Class A), Glutamate (Class C), Adhesion (Class B2), Frizzled/taste2 (Class F), and Secretin (Class B) (Bjarnadóttir et al. Reference Bjarnadóttir, Gloriam, Hellstrand, Kristiansson, Fredriksson and Schiöth2006). In addition to these major families, some organisms have lineage-specific receptors that establish distinct GPCR families (Hofmann and Palczewski Reference Hofmann and Palczewski2015). Rhodopsin-like receptors (Class A) are the most common of all known GPCRs. They are distinguished by short N-termini and their ability to interact with a wide range of ligands. The Glutamate receptor family is characterized by long N-termini that function as the binding site for ligands. Similarly, Adhesion receptors possess long N-termini containing a variety of domains, whereas Frizzled receptors feature long, cysteine-rich N-termini (Lagerström and Schiöth Reference Lagerström and Schiöth2008).

All major GPCR subfamilies were represented in schistosomes, and most of them respond to classical biogenic amines and neurotransmitters like dopamine, histamine, and serotonin (El-Shehabi et al. Reference El-Shehabi, Taman, Moali, El-Sakkary and Ribeiro2012; Hahnel et al. Reference Hahnel, Quack, Parker-Manuel, Lu, Vanderstraete, Morel, Dissous, Cailliau and Grevelding2014; MacDonald et al. Reference MacDonald, Kimber, Day and Ribeiro2015; Patocka et al. Reference Patocka, Sharma, Rashid and Ribeiro2014; Ribeiro et al. Reference Ribeiro, Gupta and El-Sakkary2012). Schistosoma GPCRs (SmGPCRs) are detected at the cell membrane and have a typical GPCR structure, an extracellular N-terminus, and an intracellular C-terminus. Most of homologues of SmGPR are characterized by the replacement of the highly conserved aspartate D3.32 of TM domain 3 with asparagine (Hamdan et al. Reference Hamdan, Abramovitz, Mousa, Xie, Durocher and Ribeiro2002). Zamanian et al. (Reference Zamanian, Kimber, McVeigh, Carlson, Maule and Day2011) identified 117 S. mansoni G PCRs genes that include all major families; 105 Rhodopsin, 2 Glutamate, 3 Adhesion, 2 Secretin, and 5 Frizzled. Among these gene receptors, novel receptor groups have been detected, including a highly diverged Platyhelminth-specific Rhodopsin and atypical Glutamate-like receptors. Genome sequencing of S. mansoni has identified 126 GPCRs (Hahnel et al. Reference Hahnel, Wheeler, Lu, Wangwiwatsin, McVeigh, Maule, Berriman, Day, Ribeiro and Grevelding2018; Kamara et al. Reference Kamara, Thao, Kaur, Wheeler and Chan2023). However, only a few of these GPCRs have been characterized in terms of its molecular and functional properties (Hoffmann et al. Reference Hoffmann, Davis, Fischer and Wynn2001; MacDonald et al. Reference MacDonald, Kimber, Day and Ribeiro2015; Patocka et al. Reference Patocka, Sharma, Rashid and Ribeiro2014; Taman and Ribeiro Reference Taman and Ribeiro2009). The diversity of GPCR genes in S. mansoni indicates a wide array of functions, potentially including reproductive development (Hahnel et al. Reference Hahnel, Wheeler, Lu, Wangwiwatsin, McVeigh, Maule, Berriman, Day, Ribeiro and Grevelding2018).

Rhodopsin receptors

Rhodopsin receptors are prototypical GPCRs (Class A). They are further characterized by a relatively short extracellular N-terminus, which is typically glycosylated, and an intracellular C-terminal tail of variable length (Kristiansen Reference Kristiansen2004). Analysis of mammalian genomes revealed that the Rhodopsin family is divided into four main groups (α, β, γ, and δ) (Fredriksson et al. Reference Fredriksson, Lagerström, Lundin and Schiöth2003). The α and β subfamilies are the only subfamilies present in S. mansoni. Alpha receptors contain amines (the largest group), opsin-like receptors, and melatonin receptors. S. mansoni possesses at least 24 putative aminergic receptors and four melanopsin-like receptors, but no melatonin-like receptors. The β subfamily contains the neuropeptide and peptide hormone GPCRs. S. mansoni contains 36 putative peptide receptors. In addition, unclassified Rhodopsin receptors have been found. A new receptor, Platyhelminth Rhodopsin Orphan Family 1, has been identified. These receptors, although displaying remnants of classical Rhodopsin, do not show homology to any previously identified GPCRs (Zamanian et al. Reference Zamanian, Kimber, McVeigh, Carlson, Maule and Day2011).

Ortholog of Rhodopsin GPCRs identified in S. mansoni miracidia share similarity with Rhodopsin GPCRs of the intermediate host B. glabrata. These GPCRs may detect similar ligands, including snail-derived odorants that could facilitate miracidial host finding (Phan et al. Reference Phan, Liang, Zhao, Wyeth, Fogarty, Duke, McManus, Wang and Cummins2022).

Rhodopsin Alpha (α) subfamily receptors

S. mansoni has histamine receptors belonging to Class A (rhodopsin-like) GPCRs. Histamine is strongly myo-excitatory in S. mansoni and is endogenously biosynthesized (Hamdan and Ribeiro Reference Hamdan and Ribeiro1999). A histamine receptor called SmGPR-1 (Smp_043260) (formerly SmGPCR) was cloned in S. mansoni. SmGPR-1 has a structure characteristic of the amine GPCR family but does not obviously resemble any of the histamine receptors in mammals. Histamine activation of SmGPCR triggered mobilization of intracellular calcium, but not cAMP. Furthermore, SmGPCR-1 showed a glycine (Gly196) substitution instead of (Asn/Thr) or charged residue (Glu) in TM domain 5 and an asparagine (Asn111) instead of aspartate of TM domain 3 (Hamdan et al. Reference Hamdan, Abramovitz, Mousa, Xie, Durocher and Ribeiro2002). El-Shehabi et al. (Reference El-Shehabi, Vermeire, Timothy and Yoshino2009) revealed that this receptor is expressed in the tegument and musculature of both cercariae and adult parasites.

SmGPR-2 (Smp_043340) is a second histamine receptor of S. mansoni. It is an orphan receptor expressed in the vicinity of histamine-containing neurons in the sub-tegumental neuronal plexus. It is developmentally regulated showing up-regulation in the parasitic stages compared to cercaria, with the highest level of expression in young schistosomula. The highly conserved aspartate D3.32 of TM domain 3 is also absent in SmGPR-2. This receptor has a novel pharmacological profile. It is inhibited by drugs not known to interact with histamine receptors, while classical anti-histamines had no effect on the receptor activity (El-Shehabi and Ribeiro Reference El-Shehabi and Ribeiro2010).

Dopamine receptors are GPCRs of the Class A Rhodopsin family. Mammals and invertebrates possess five dopamine receptors (D1-D5), which are categorized into two classes, D1-type and D2-type, based on their amino acid sequence homology and pharmacological profiles. D1-type dopamine receptors (D1 and D5) are associated with G-stimulatory proteins. Its activation leads to stimulation of adenylyl cyclase, resulting in an increase in cyclic adenosine monophosphate production from adenosine triphosphate (Gurevich et al. Reference Gurevich, Gainetdinov and Gurevich2016). In contrast, D2-type dopamine receptors (D2, D3, and D4) are linked to G-inhibitory proteins, which inhibit adenylyl cyclase and decrease cyclic adenosine monophosphate levels. S. mansoni D2 (SmD2) (Smp_127310) dopamine receptor exhibits an unusual pharmacological profile. Apomorphine, a potent antagonist of mammalian D2-type receptors, acts as an agonist for the SmD2 receptor, while other classic mammalian antagonists have no effect. This receptor is found in the membrane protein fractions of S. mansoni cercaria, schistosomula, and adult worms. SmD2 is also present in the sub-tegumental somatic musculature and acetabulum of cercaria and schistosomula. In adult parasites, SmD2 is enriched in the somatic muscles and, to a lesser extent, in the muscular lining of the caecum (Taman and Ribeiro Reference Taman and Ribeiro2009). In miracidium, antagonists of D2-type receptors have been found to delay miracidial transformation (Taft et al. Reference Taft, Norante and Yoshino2010).

Another neurotransmitter dopaminergic receptor belonging to GPCRs has been identified in S. mansoni and is named SmGPR-3 (Smp_043290) (El-Shehabi et al. Reference El-Shehabi, Taman, Moali, El-Sakkary and Ribeiro2012). This receptor is an orphan amine-like receptor found in schistosomes but not in mammals and has an atypical antagonist profile compared to mammalian receptors. Some mammalian D2 antagonists enhanced the activity of SmGPR-3 (El-Shehabi et al. Reference El-Shehabi, Taman, Moali, El-Sakkary and Ribeiro2012). SmGPR-3 is abundantly expressed in the nervous system of schistosomes, particularly in the main nerve cords and in the peripheral innervation of body wall muscles. Therefore, there are at least two routes of dopaminergic motor control in S. mansoni, involving both direct and indirect mechanisms. One pathway is mediated by SmD2, which is predicted to act directly on the musculature, while the other is a more indirect neuronal pathway mediated by SmGPR-3 (El-Shehabi et al. Reference El-Shehabi, Taman, Moali, El-Sakkary and Ribeiro2012).

In addition to histamine and dopamine, serotonin is one of the best characterized amines in flatworms and causes muscle excitation in S. mansoni. The worm contains a serotonergic receptor (Sm5HTR) that belongs to the Class A Rhodopsin family and is distantly related to serotonergic type 7 (5HT7) receptors found in other species. Sm5HTR signals through an increase in intracellular cAMP. The receptor is distributed in the cerebral ganglia and main nerve cords and in peripheral nerves of the body wall muscles and tegument. The serotonin receptor (Smp_126730) is a crucial component of the motor control system in S. mansoni (Patocka et al. Reference Patocka, Sharma, Rashid and Ribeiro2014). It is worth mentioning that PZQ, the sole treatment of schistosomiasis, has been identified as a GPCR ligand that acts by modulating serotoninergic signalling. PZQ modulates serotonergic signalling within a concentration range adequate to regulate the vascular tone of mesenteric blood vessels, where adult parasites reside in their host. The activity of PZQ on both parasite and host GPCRs likely contributes to its clinical efficacy by combining a harmful paralytic effect on the parasite with favorable effects on the host that aid in worm clearance (Chan et al. Reference Chan, Cupit, Gunaratne, McCorvy, Yang, Stoltz, Webb, Dosa, Roth, Abagyan, Cunningham and Marchant2017).

Octopamine and its precursor tyramine (phenolamines) are invertebrate specific biogenic proteins and neurotransmitter derived from tyrosine. They are considered the invertebrate counterpart of the adrenergic system. The schistosome genome annotated two putative GPCRs; octopamine (Smp_150180) and tyramine (Smp_043290) GPCRs (Protasio et al. Reference Protasio, Tsai, Babbage, Nichol, Hunt, Aslett, De Silva, Velarde, Anderson, Clark, Davidson, Dillon, Holroyd, LoVerde, Lloyd, McQuillan, Oliveira, Otto, Parker-Manuel, Quail, Wilson, Zerlotini, Dunne and Berriman2012). Octopamine receptor is a G protein-coupled receptor (GPCR) belonging to class A Rhodopsin-like subfamily (Hill et al. Reference Hill, Sharan and Watts2018). Octopamine labeling leads to the discovery of two pairs of ganglia in the adult schistosome brain. This neurotransmitter is localized in both ganglia and is also distributed throughout central and peripheral nerves and modulates schistosomula motility and length (El-Sakkary et al. Reference El-Sakkary, Chen, Arkin, Caffrey and Ribeiro2018).

Rhodopsin Beta (β) subfamily receptors

The β subfamily contains the great majority of neuropeptide and neuropeptide hormone GPCRs. Their neuropeptide signalling is known to play an essential role in flatworm locomotion, feeding, reproduction, host-finding, and regeneration (Kreshchenko Reference Kreshchenko2008). The genome of S. mansoni identifies at least 14 potential neuropeptide receptors, including several FLP-like and NPY/F-like receptors (Berriman et al. Reference Berriman, Haas, LoVerde, Wilson, Dillon, Cerqueira, Mashiyama, Al-Lazikani, Andrade, Ashton, Aslett, Bartholomeu, Blandin, Caffrey, Coghlan, Coulson, Day, Delcher, DeMarco, Djikeng, Eyre, Gamble, Ghedin, Gu, Hertz-Fowler, Hirai, Hirai, Houston, Ivens, Johnston, Lacerda, Macedo, McVeigh, Ning, Oliveira, Overington, Parkhill, Pertea, Pierce, Protasio, Quail, Rajandream, Rogers, Sajid, Salzberg, Stanke, Tivey, White, Williams, Wortman, Wu, Zamanian, Zerlotini, Fraser-Liggett, Barrell and El-Sayed2009). The invertebrate neuropeptide F family is related to the neuropeptide Y family of vertebrate peptides (with a C-terminal F instead of a Y) (McVeigh et al. Reference McVeigh, Mair, Atkinson, Ladurner, Zamanian, Novozhilova, Marks, Day and Maule2009). Zamanian et al. (Reference Zamanian, Kimber, McVeigh, Carlson, Maule and Day2011) identified several peptide receptors in S. mansoni, denoting that the peptidergic signalling is important for neurotransmission in the worm. They found that the number of potential flatworm peptide receptors significantly exceeds the peptide ligands identified so far. Most of the identified receptors cannot be associated with specific ligands with certainty. Peptides include FMRF amide-like peptides (FLPs), neuropeptide Fs (NPFs), and various other specific amides, some of which have similarities to peptides found in other phyla, such as neuropeptide FF (NPFF)-like and gonadotropin- or thyrotropin-releasing hormone-like peptides.

S. mansoni neuropeptide Y/F and its receptors have been identified in intramolluscan stages of S. mansoni and have been found to be associated with maintenance of schistosome germinal cell during intramolluscan development (Buddenborg et al. Reference Buddenborg, Kamel, Hanelt, Bu, Zhang, Mkoji and Loker2019). This neuropeptide has been also linked to a reduction in egg production of the infected snails (de Jong-Brink et al. Reference de Jong-Brink, ter Maat and Tensen2001).

Rhodopsin orphan receptor

A transcriptomics study revealed a rhodopsin orphan GPCR20 of S. mansoni (SmGPCR20) (Li et al. Reference Li, Weth, Haimann, Möscheid, Huber and Grevelding2024). This receptor belongs to the paired males-unpaired males-unpaired females subgroup of SmGPCRs, which is differentially transcribed between males and females expressing high transcript levels in paired male worms (bM), unpaired male worms (sM), and unpaired females (sF), whereas low or no transcripts of this subgroup were present in paired female worms (bF) (Hahnel et al. Reference Hahnel, Wheeler, Lu, Wangwiwatsin, McVeigh, Maule, Berriman, Day, Ribeiro and Grevelding2018; Lu et al. Reference Lu, Sessler, Holroyd, Hahnel, Quack, Berriman and Grevelding2016). This SmGPCR20 orphan receptor (Smp_084270) interacts with two neuropeptides – SmNPP26 and SmNPP40 – as potential interaction legends. qRT-PCR revealed that Smgpcr20, Smnpp26, and Smnpp40 genes showed sex- and/or pairing-dependent expression. The combination of SmGPCR20 with these neuropeptides affects egg production, oogenesis, and growth of the S. mansoni females (Li et al. Reference Li, Weth, Haimann, Möscheid, Huber and Grevelding2024).

Adhesion and Secretin receptors

Adhesion and Secretin receptors belong to Class B2 and Class B, respectively. In vertebrates, this family represents the second largest group of GPCRs, following the Rhodopsin family. Adhesion and Secretin receptors were identified in S. mansoni (Zamanian et al. Reference Zamanian, Kimber, McVeigh, Carlson, Maule and Day2011). They share sequence similarity in their 7-TM domains, but they showed structural differences in their N-terminal domains. Adhesion GPCRs possess a long N-terminal domain that features a varied arrangement of functional domains. Secretin GPCRs have N-terminal hormone-binding domains (HBD) that enable them to respond to peptide hormones (Nordström et al. Reference Nordström, Lagerström, Wallér, Fredriksson and Schiöth2009).

Cellular adhesion molecules are involved in the pathogenesis of S. mansoni. Integrins are one of the cellular adhesion molecules that also include cadherins, selectins, and the immunoglobulin (Ig) (Figliuolo et al. Reference Figliuolo da Paz, Figueiredo-Vanzan and Dos Santos Pyrrho2019). Integrins are a family of heterodimeric transmembrane receptors comprising at least 18 α and 8 β subunits in mammals (Hynes Reference Hynes2002). They mediate cell adhesion, functioning as connectors between the extracellular matrix and the cytoskeleton, while transmitting biochemical and mechanical signals between cells and their surroundings. They function bidirectionally across the plasma membrane, facilitating both inside-out and outside-in signalling (Fu et al. Reference Fu, Wang and Luo2012). Integrins are also involved in various immune-related signalling pathways in mammals (Zhang et al. Reference Zhang, Zhang, Chen and Xie2023).

Integrins work synergistically with other molecules, such as VKR1 and RTKs, within complex signalling pathways that regulate growth and differentiation processes (Gelmedin et al. Reference Gelmedin, Morel, Hahnel, Cailliau, Dissous and Grevelding2017). A significant portion of integrin signalling functions is dependent on a cytoplasmic TK (Harburger and Calderwood Reference Harburger and Calderwood2009). In S. mansoni, the genes coding for these signalling transduction protein kinases have been shown to have roles in reproductive organs differentiation process (Knobloch et al. Reference Knobloch, Beckmann, Burmeister, Quack and Grevelding2007). These kinases are cellular tyrosine kinases members that act in a multi-kinase complex (Beckmann et al. Reference Beckmann, Hahnel, Cailliau, Vanderstraete, Browaeys, Dissous and Grevelding2011) such as Src (SmTK3) (Kapp et al. Reference Kapp, Knobloch, Schüssler, Sroka, Lammers, Kunz and Grevelding2004), Syk (SmTK4) (Beckmann et al. Reference Beckmann, Buro, Dissous, Hirzmann and Grevelding2010), and Src/Abl (SmTK6) families (Beckmann et al. Reference Beckmann, Hahnel, Cailliau, Vanderstraete, Browaeys, Dissous and Grevelding2011).

Beckmann et al. (Reference Beckmann, Quack, Dissous, Cailliau, Lang and Grevelding2012) characterized four alpha-integrins (Smα-Int1–Smα-Int4) and one beta-integrin (Smβ-Int1) subunit from S. mansoni. The α (Smα-Int1, Smp_126140 and Smα-Int2, Smp_170280) and β (Smβ-Int1, Smp_089700) subunits are also present in S. mansoni exosomes (Samoil et al. Reference Samoil, Dagenais, Ganapathy, Aldridge, Glebov, Jardim and Ribeiro2018). The α-integrins of the free-living planarian Schmidtea mediterranea differ from those of S. mansoni in having only three α-integrin subunits. The β-integrins of S. mansoni did not bind fibronectin (Beckmann et al. Reference Beckmann, Quack, Dissous, Cailliau, Lang and Grevelding2012). These β-integrins are closely related to the human β4 subunit that binds laminin (Hynes Reference Hynes2002). The Smβ-Int1/Smα-Int1 heterodimer are found in the gonads, whereas the Smβ-Int1/Smα-Int2 heterodimer might fulfill more specialized functions in the area surrounding the ootype. Smβ-Int1 interacts and co-localizes with cellular tyrosine kinases in the reproductive organs of schistosomes with SmTK4, a Syk kinase, being its most significant interaction partner (Beckmann et al. Reference Beckmann, Quack, Dissous, Cailliau, Lang and Grevelding2012). The Smβ-Int1/SmVKR1 signalling complex plays a crucial role in the oocyte differentiation and survival in paired schistosomes (Gelmedin et al. Reference Gelmedin, Morel, Hahnel, Cailliau, Dissous and Grevelding2017). If Schistosoma integrins function in the same way as the mammalian homologues, these proteins could also help the parasite modulate the immune response of host cells (Samoil et al. Reference Samoil, Dagenais, Ganapathy, Aldridge, Glebov, Jardim and Ribeiro2018).

Glutamate receptors (GluRs)

L-glutamate is an important amino acid neurotransmitter in vertebrates and many invertebrates. It exerts its effects through interactions with ionotropic glutamate receptors (iGluRs) and metabotropic glutamate receptors (mGluRs). The iGluRs function as voltage-gated ion channels, while the metabotropic glutamate receptors (mGluRs) belong to the Class C GPCR superfamily, characterized by the typical seven transmembrane domain structure (Reiner and Levitz Reference Reiner and Levitz2018). Glutamate-gated chloride channels (GluCls) (Smp_128940) are pentameric ligand-gated inhibitory ion channels found exclusively in invertebrates. Their absence in vertebrates makes them an ideal target for antiparasitic drugs. However, GluCls of S. mansoni worms differ significantly from the GluCls of nematodes. This is exemplified by ivermectin, which leads to flaccid paralysis or kills roundworms by activating GluCls, while schistosomes are not susceptible to the drug (Dufour et al. Reference Dufour, Beech, Wever, Dent and Geary2013). Callau-Vázquez et al. (Reference Callau-Vázquez, Pless and Lynagh2018) demonstrated that the GluCl-2 from S. mansoni is activated by glutamate with a potency similar to that of nematode GluCls, despite substantial divergence in the ligand-binding C loop that differs in length compared to other pentameric ligand-gated ion channels, as well as the difference in hydrophobic channel gate.

Upon binding glutamate, mGluRs trigger signalling cascades or facilitate cation influx. mGluRs are structurally related to metabotropic gamma-aminobutyric acid (GABA) receptors and calcium-sensing, taste, and pheromone receptors (Niswender and Conn Reference Niswender and Conn2010). mGluRs are divided into three main groups. Group I (mGluR1 and mGluR5) signals through changes in intracellular calcium and the inositol phospholipid pathway, whereas Group II (mGluR2 and mGluR3) and Group III (mGluR4, mGluR6, mGluR7 and mGluR8) signal primarily through inhibition of adenylate cyclase, which in turn decreases intracellular cAMP signalling (Kryszkowski and Boczek Reference Kryszkowski and Boczek2021). mGluRs consist of a large N-terminal extracellular ligand binding domain (LBD), the characteristic 7-transmembrane (7-TM) segment, and a variable-length intracellular C-terminal domain (ICD). The LBD of mGluRs contains the glutamate-binding site within a Venus Flytrap module, which is connected to the 7-TM region by a short cysteine-rich domain (CRD) (Niswender and Conn Reference Niswender and Conn2010) (Figure 3a).

Figure 3. (a) Schematic diagram of the expected S. mansoni metabotropic glutamate receptor (SmGluR) (Smp_128940) comprising a large N-terminal extracellular ligand binding domain (LBD), a seven-transmembrane (7-TM) anchoring segment, and a C-terminal intracellular domain (ICD) of varying lengths. The LBD is connected to the 7-TM region by a short cysteine-rich domain (CRD). (b) S. mansoni glutamate-binding protein (SmGBP) (Smp_052660) receptor has a conserved ligand binding domain (LBD) but is missing the cysteine-rich domain, the characteristic 7-TM region, and intracellular domain.

Glutamate immunoreactivity in S. mansoni was detected in the nervous system, including the cerebral ganglia, longitudinal nerve cords, and commissures (Mendonça-Silva et al. Reference Mendonça-Silva, Pessôa and Noël2002). The genome of S. mansoni encodes at least three sequences that are homologous to mGluRs from other species (Berriman et al. Reference Berriman, Haas, LoVerde, Wilson, Dillon, Cerqueira, Mashiyama, Al-Lazikani, Andrade, Ashton, Aslett, Bartholomeu, Blandin, Caffrey, Coghlan, Coulson, Day, Delcher, DeMarco, Djikeng, Eyre, Gamble, Ghedin, Gu, Hertz-Fowler, Hirai, Hirai, Houston, Ivens, Johnston, Lacerda, Macedo, McVeigh, Ning, Oliveira, Overington, Parkhill, Pertea, Pierce, Protasio, Quail, Rajandream, Rogers, Sajid, Salzberg, Stanke, Tivey, White, Williams, Wortman, Wu, Zamanian, Zerlotini, Fraser-Liggett, Barrell and El-Sayed2009). Taman and Ribeiro (Reference Taman and Ribeiro2011a) described a mGluR in S. mansoni (SmGluR) in the nervous system of adult worms and cercariae, as well as in the female reproductive tract. SmGluR belongs to the GPCR superfamily and shares a distant relationship with mGluRs found in other species (Figure 3a). However, SmGluR differs from mammalian mGluRs with respect to signalling mechanism and pharmacological profile. SmGluR is activated by glutamate, whereas GABA has no significant effect. Phylogenetic analyses indicated that SmGluR shares a similar degree of sequence homology with mGluRs as it does with other family C GPCRs, such as GABA receptors.

The second metabotropic glutamate receptor identified in S. mansoni is glutamate-binding protein (SmGBP) (Smp_052660) (Taman and Ribeiro Reference Taman and Ribeiro2011b). SmGBP represents a new type of glutamate receptor that may be unique to flatworms. Genes encoding similarly truncated receptors have been found in the S. japonicum genome (The Schistosoma japonicum Genome Sequencing and Functional Analysis Consortium 2009) and the partially annotated genome of the planarian Schmidtea mediterranea (Robb et al. Reference Robb, Ross and Sánchez Alvarado2008), but they are not known to occur in other metazoans. SmGBP receptor is an atypical receptor; it is a C-terminally truncated mGluR with a conserved ligand (glutamate)-binding domain (LBD) located within a Venus Flytrap module but lacking the cysteine-rich domain (CRD), the characteristic 7-TM region, and the intracellular domain (ICD) (Figure 3b). SmGBP is suggested to be either an integral membrane protein or a peripheral protein closely associated with the membrane. This receptor is gender- and stage-specific. SmGBP is localized on the surface of male worms, especially on the dorsal tubercles but not in females or larval stages (Taman and Ribeiro Reference Taman and Ribeiro2011b). In S. japonicum, the related schistosome species, two putative mGluRs, are identified: SjGRM7 and SjGRM. SjGRM7 has been found to be crucial for normal physiological functions, growth, development, and egg production (Wang et al. Reference Wang, Cheng, Chen, Zhang, Xie, Liu, You, Yi, Zhu, Gu, Xu, Lu, Wang and Hu2022).

Frizzled receptors

Frizzled protein receptors belong to GPCRs Class F. They consist of seven trans-membrane proteins with a cysteine-rich domain in the N-terminal extracellular region required for Wnt ligand binding. Most Frizzled receptors share a common C-terminal motif that is a binding site for the cytoplasmic protein domain (Hering and Sheng Reference Hering and Sheng2002). Wnt signalling plays a key role in embryonic development, energy metabolism, and balance (Nusse Reference Nusse2015). Zamanian et al. (Reference Zamanian, Kimber, McVeigh, Carlson, Maule and Day2011) identified four Frizzled sequences S. mansoni. SmFz1 genes (Smp_118970, Smp_173940) are regulated by pairing in gonads. In vitro inhibition of the gene affects the survival of adult worms, decreases the egg production, and affects the gonad differentiation, morphology, and embryogenesis (Hahnel et al. Reference Hahnel, Quack, Parker-Manuel, Lu, Vanderstraete, Morel, Dissous, Cailliau and Grevelding2014). Secreted frizzled-related protein can inhibit Wnt signalling by competitive binding to the frizzled protein-specific receptor (García-Tobilla et al. Reference García-Tobilla, Solórzano, Salido-Guadarrama, González-Covarrubias, Morales-Montor, Díaz-Otañez and Rodríguez-Dorantes2016). Knockdown of S. japonicum secreted frizzled-related protein gene impairs worm growth and development, survival and morphological structure, reproductive ability, and viability of the eggs produced (Cheng et al. Reference Cheng, Li, Qin, Xu, Zhang, Liu, Gu and Jin2019).

Acetylcholine receptors

Acetylcholine (ACh) is a crucial neurotransmitter in both vertebrate and invertebrate species. In vertebrates, ACh functions primarily as an excitatory neurotransmitter, regulating processes such as muscle contraction, glandular secretion, and memory formation. ACh similarly plays an excitatory role in invertebrates, and its involvement in nematode motor function is well documented. However, there is a significant exception in schistosomes, where ACh acts as a major inhibitory neurotransmitter or modulator. Activation of ACh receptors (AChR) in S. mansoni leads to muscle relaxation, resulting in flaccid paralysis (Day et al. Reference Day, Chen, Miller, Tian, Bennett and Pax1996). Metrifonate, an Acetylcholinesterase (AChE) inhibitor, elevates synaptic levels of ACh, resulting in prolonged paralysis of the axial muscles of schistosomes and halting its movement. This action is suggested to be due to the secondary effects of muscle paralysis. The drug demonstrates equal potency and efficacy in vitro against both S. mansoni and S. haematobium but is only effective in vivo against the latter species (Bueding et al. Reference Bueding, Liu and Rogers1972).

Most of AChRs in schistosomes are nicotinic AChRs (nAChRs), so named because of their high affinity for nicotine. However, muscarinic cholinergic receptors are also expected to be present. One of these receptors possesses all the structural characteristics of GPCR (MacDonald et al. Reference MacDonald, Kimber, Day and Ribeiro2015). In vertebrates, nAChRs are invariably cation-selective (Na+, Ca2+, K+) and mediate excitatory responses. In contrast, invertebrates have cation and anion-selective (Cl) ACh-gated channels. These acetylcholine-gated chloride channels (ACC) mediate Cl--driven membrane hyperpolarization and are believed to play a role in inhibitory responses to ACh. These ACC are structurally related to nAChRs but are selective for chloride ions (Beech et al. Reference Beech, Callanan, Rao, Dawe and Forrester2013). Structurally, nAChRs belong to the superfamily of Cys-loop ligand-gated ion channel. They form homo- and hetero-pentameric structures organized in a barrel shape around a central ion-selective pore (Albuquerque et al. Reference Albuquerque, Pereira, Alkondon and Rogers2009). A key characteristic of ACCs is the presence of a Pro-Ala motif in the pore-lining M2 domains of their subunits. This motif has been shown to convey anion selectivity to other ligand-gated ion channels (LGICs), replacing a Glu residue typically found in cation-selective channels (Keramidas et al. Reference Keramidas, Moorhouse, Pierce, Schofield and Barry2002). These ACCs that appear to be specific to invertebrates are found in S. mansoni (SmACCs) and have an inhibitory modulatory effect on the neuromuscular system of schistosome potentially through a chloride influx produced by the activation of SmACCs and their receptors (SmACC-1 and SmACC-2) (Smp_176310 and Smp_142690) (MacDonald et al. Reference MacDonald, Buxton, Kimber, Day, Robertson and Ribeiro2014). Treatment with ACh antagonists and RNA interference (RNAi) leads to suppression of SmACCs and induces a hypermotile effect. Two of the SmACCs were localized to regions of the peripheral nervous system that innervate the body wall muscles; however, none appear to be directly expressed in the muscle tissue (MacDonald et al. Reference MacDonald, Buxton, Kimber, Day, Robertson and Ribeiro2014).

Muscarinic acetylcholine receptors (mAChRs) belong to GPCR superfamily and are related to Rhodopsin (Family A GPCRs) in their structure. The term ‘muscarinic’ originates from the preference of receptors to bind to and be activated by the fungal toxin muscarine (Dale Reference Dale1914). Schistosome muscarinic acetylcholine receptor is also referred to as G protein-coupled acetylcholine receptors (SmGAR) (Smp_145540). Expression of this receptor is predicted to be high during the early larval stages of schistosomes (Protasio et al. Reference Protasio, Tsai, Babbage, Nichol, Hunt, Aslett, De Silva, Velarde, Anderson, Clark, Davidson, Dillon, Holroyd, LoVerde, Lloyd, McQuillan, Oliveira, Otto, Parker-Manuel, Quail, Wilson, Zerlotini, Dunne and Berriman2012). SmGAR is constitutively active but can be further stimulated by ACh and, to a lesser extent, by the cholinergic agonist carbachol. Anti-cholinergic drugs exhibit an inverse agonist activity towards SmGAR, significantly reducing its basal activity. A phenotypic RNAi assay demonstrated that suppression of SmGAR activity in early-stage larval schistosomula results in a marked decrease in larval motility (MacDonald et al. Reference MacDonald, Kimber, Day and Ribeiro2015).

In addition to its neuromuscular effects, ACh has been linked to glucose transport across the tegument and increased glucose uptake in schistosomes. AChE has been shown to play a role in modulating glucose uptake by schistosomes from the blood of mammalian hosts. Two main molecular forms of AChE are found in S. mansoni. One form is located within the muscle and plays a role in cholinergic processes, while the other form is found on the surface, anchored to the membrane by a covalently bound glycophosphatidylinositol anchor. This surface-localized AChE may participate in non-cholinergic processes and signal transduction (Espinoza et al. Reference Espinoza, Silman, Arnon and Tarrab-Hazdai1991). Glycophosphatidylinositol-anchored AChE can be released from the schistosome surface membrane by a PI-specific phospholipase C, which can remove significant amounts of AChE from the tegument of schistosomula in vitro without affecting the parasite viability (Espinoza et al. Reference Espinoza, Tarrab-Hazdai, Silman and Arnon1988). It has been suggested that release of AChE triggers immediate replenishment of the surface enzyme. However, this process occurs with another glycophosphatidylinositol-anchored protein, alkaline phosphatase, which is also present on the surface of schistosome (Arnon et al. Reference Arnon, Silman and Tarrab-Hazdai1999).

Glucose uptake is regulated through the interaction of ACh with tegumental nAChRs and AChE. The effect of ACh on glucose uptake can be inhibited by blocking any of the ACh cholinergic systems. AChE is thought to regulate interaction of ACh with its receptor since inhibition of AChE produces an effect similar to excessive presence of ligand (Jones et al. Reference Jones, Bentley, Oliveros Parra and Agnew2002). Exposure to the same concentration of ACh present in host blood was found to enhance glucose uptake in S. haematobium and S. bovis, but not in S. mansoni. However, at higher concentrations, ACh inhibited glucose uptake from the host blood into the parasites. The glucose uptake rate in adult S. haematobium and S. bovis is roughly double that of S. mansoni (Camacho and Agnew Reference Camacho and Agnew1995), and the first two species have relatively higher AChE activity on their teguments compared to S. mansoni (Camacho et al. Reference Camacho, Tarrab-Hazdai, Espinoza, Arnon and Agnew1994). These elevated levels of AChE activity contribute to its increased susceptibility to metrifonate (Harder Reference Harder2002), which may explain why metrifonate is effective against S. haematobium and S. bovis but not S. mansoni.

Adult stages of schistosomes possess AChE and nAChR on their teguments, and both components are concentrated on the surface of the adult male, a key site for nutrient uptake for the worm pair (Camacho and Agnew Reference Camacho and Agnew1995). AChR expression increases during parasites pairing and sexual maturation as the pairing state increases the uptake of several host compounds (Camacho et al. Reference Camacho, Tarrab-Hazdai, Espinoza, Arnon and Agnew1994). AChE inhibitors impair the parasite glucose uptake ability, which affects the parasite growth and development (Sundaraneedi et al. Reference Sundaraneedi, Tedla, Eichenberger, Becker, Pickering, Smout, Rajan, Wangchuk, Keene, Loukas, Collins and Pearson2017; You et al. Reference You, Liu, Du, Nawaratna, Rivera, Harvie, Jones and McManus2018).

Conclusion and perspectives

The success of Schistosoma mansoni infections is partly attributed to its ability to utilize host-derived molecules through several receptors, which are essential for its growth and development. These receptors play a coordinated role in regulating the parasite’s life processes, using growth factors from both the parasite itself and host-derived molecules. The key receptors involved include growth factor receptors, nuclear hormone receptors, nuclear transport receptors, and neurotransmitter receptors. A deeper understanding of how schistosomes exploit host nutrients, neuro-endocrine hormones, and signalling pathways for their growth, development, and maturation is expected to lead to improved interventions to control schistosomiasis.

More and more new receptors, along with related proteins, ligands, and genes, are being identified and characterized in schistosomes, especially with the availability of extensive genomic data for S. mansoni. Understanding the molecular roles that these receptors play in S. mansoni growth, as well as developing more specific receptor agonists and antagonists, presents a major challenge for future research. Notably, many of these receptors share minimal sequence homology with those of the human host, making them particularly suitable for selective drug targeting.

Acknowledgements

Not applicable.

Author contribution

Iman Abou-El-Naga contributed to the study conception and design. Material preparation and data collection were performed by Iman Abou-El-Naga. The manuscript was written and approved by Iman Abou-El-Naga.

Financial support

This study has not received any funding.

Competing interest

The author declares none.

Ethical standard

The protocol of the present study was approved by the ethics Committee of the Faculty of Medicine, Alexandria University according to the institutional ethical guidelines.

References

Abou-El-Naga, IF and Radwan, EH (2012) Defense response of susceptible and resistant Biomphalaria alexandrina snails against Schistosoma mansoni infection. Revista de Biologia Tropical 60, 11951204. https://doi.org/10.15517/rbt.v60i3.1771.Google ScholarPubMed
Abou-El-Naga, IF, Sadaka, HA, Amer, EI, Diab, IH and Khedr, SI (2015) Impact of the age of Biomphalaria alexandrina snails on Schistosoma mansoni transmission: Modulation of the genetic outcome and the internal defence system of the snail. Memórias do Instituto Oswaldo Cruz 110, 585595. https://doi.org/10.1590/0074-02760150016.CrossRefGoogle ScholarPubMed
Abou-El-Naga, IF (2015) Demographic, socioeconomic and environmental changes affecting circulation of neglected tropical diseases in Egypt. Asian Pacific Journal of Tropical Medicine 8, 881888. https://doi.org/10.1016/j.apjtm.2015.10.015.CrossRefGoogle ScholarPubMed
Abou-El-Naga, IF (2018) Towards elimination of schistosomiasis after 5000 years of endemicity in Egypt. Acta Tropica 181, 112121. https://doi.org/10.1016/j.actatropica.2018.02.005.CrossRefGoogle ScholarPubMed
Abou-El-Naga, IF (2021) Review: Schistosoma mansoni phosphatidylinositol 3 kinase (PI3K)/Akt/mechanistic target of rapamycin (mTOR) signaling pathway. Comparative Biochemistry and Physiology Part B Biochemistry and Molecular Biolology 256, 110632. https://doi.org/10.1016/j.cbpb.2021.110632.CrossRefGoogle ScholarPubMed
Abreu, FC, Pereira, RV, Oliveira, VF, Gomes Mde, S, Jannotti-Passos, LK, Borges, WC and Guerra-Sá, R (2013) Characterization of export receptor exportins (XPOs) in the parasite Schistosoma mansoni. Parasitology Research 112, 41514159. https://doi.org/10.1007/s00436-013-3606-x.CrossRefGoogle ScholarPubMed
Ahier, A, Khayath, N, Vicogne, J and Dissous, C (2008) Insulin receptors and glucose uptake in the human parasite Schistosoma mansoni. Parasite 15, 573579. https://doi.org/10.1051/parasite/2008154573.CrossRefGoogle ScholarPubMed
Albuquerque, EX, Pereira, EF, Alkondon, M and Rogers, SW (2009) Mammalian nicotinic acetylcholine receptors: From structure to function. Physiological Reviews 89, 73120. https://doi.org/10.1152/physrev.00015.2008.CrossRefGoogle ScholarPubMed
AlHariry, NS, El Saftawy, EA, Aboulhoda, BE, Abozamel, AH, Alghamdi, MA, Hamoud, AE and Khalil Ghanam, WAE (2024) Comparison of tissue biomarkers between non-schistosoma and schistosoma-associated urothelial carcinoma. Tissue and Cell 88, 102416. https://doi.org/10.1016/j.tice.2024.102416.CrossRefGoogle ScholarPubMed
Amer, EI, Abou-El-Naga, IF, Boulos, LM, Ramadan, HS and Younis, SS (2022) Praziquantel-encapsulated niosomes against Schistosoma mansoni with reduced sensitivity to praziquantel. Biomedica 42, 6784. https://doi.org/10.7705/biomedica.5913.CrossRefGoogle ScholarPubMed
Andrade, LF, Nahum, LA, Avelar, LG, Silva, LL, Zerlotini, A, Ruiz, JC and Oliveira, G (2011) Eukaryotic protein kinases (ePKs) of the helminth parasite Schistosoma mansoni. BMC Genomics 12, 215234. https://doi.org/10.1186/1471-2164-12-215.CrossRefGoogle ScholarPubMed
Arnon, R, Silman, I and Tarrab-Hazdai, R (1999) Acetylcholinesterase of Schistosoma mansoni: functional correlates. Protein Science 8, 25532561. https://doi.org/10.1110/ps.8.12.2553.CrossRefGoogle ScholarPubMed
Avelar, LG, Nahu, LA, Andrade, LF and Oliveira, G (2011) Functional diversity of the Schistosoma mansoni tyrosine kinases. Journal of Signal Transduction 2011, 603290. https://doi.org/10.1155/2011/603290.CrossRefGoogle ScholarPubMed
Baba, AB, Rah, B, Bhat, GR, Mushtaq, I, Parveen, S, Hassan, R, Hameed Zargar, M and Afroze, D (2022) Transforming growth factor-beta (TGF-β) signaling in cancer – a betrayal within. Frontiers in Pharmacology 13, 791272. https://doi.org/10.3389/fphar.2022.791272.CrossRefGoogle Scholar
Beckmann, S, Buro, C, Dissous, C, Hirzmann, J and Grevelding, CG (2010) The Syk kinase SmTK4 of Schistosoma mansoni is involved in the regulation of spermatogenesis and oogenesis. PLOS Pathogens 6(2), e1000769. https://doi.org/10.1371/journal.ppat.1000769.CrossRefGoogle ScholarPubMed
Beckmann, S, Hahnel, S, Cailliau, K, Vanderstraete, M, Browaeys, E, Dissous, C and Grevelding, CG (2011) Characterization of the Src/Abl hybrid kinase SmTK6 of Schistosoma mansoni. Journal of Biological Chemistry 286, 4232542336. https://doi.org/10.1074/jbc.M110.210336.CrossRefGoogle ScholarPubMed
Beckmann, S, Quack, T, Dissous, C, Cailliau, K, Lang, G and Grevelding, CG (2012) Discovery of platyhelminth-specific α/β-integrin families and evidence for their role in reproduction in Schistosoma mansoni. PLOS One 7(12), e52519. https://doi.org/10.1371/journal.pone.0052519.CrossRefGoogle ScholarPubMed
Beech, RN, Callanan, MK, Rao, VT, Dawe, GB and Forrester, SG (2013) Characterization of Cys-loop receptor genes involved in inhibitory amine neurotransmission in parasitic and free-living nematodes. Parasitology International 62, 599605. https://doi.org/10.1016/j.parint.2013.03.010.CrossRefGoogle ScholarPubMed
Berriman, M, Haas, BJ, LoVerde, PT, Wilson, RA, Dillon, GP, Cerqueira, GC, Mashiyama, ST, Al-Lazikani, B, Andrade, LF, Ashton, PD, Aslett, MA, Bartholomeu, DC, Blandin, G, Caffrey, CR, Coghlan, A, Coulson, R, Day, TA, Delcher, A, DeMarco, R, Djikeng, A, Eyre, T, Gamble, JA, Ghedin, E, Gu, Y, Hertz-Fowler, C, Hirai, H, Hirai, Y, Houston, R, Ivens, A, Johnston, DA, Lacerda, D, Macedo, CD, McVeigh, P, Ning, Z, Oliveira, G, Overington, JP, Parkhill, J, Pertea, M, Pierce, RJ, Protasio, AV, Quail, MA, Rajandream, MA, Rogers, J, Sajid, M, Salzberg, SL, Stanke, M, Tivey, AR, White, O, Williams, DL, Wortman, J, Wu, W, Zamanian, M, Zerlotini, A, Fraser-Liggett, CM, Barrell, BG and El-Sayed, NM (2009) The genome of the blood fluke Schistosoma mansoni. Nature 460, 352358. https://doi.org/10.1038/nature08160.CrossRefGoogle ScholarPubMed
Bertin, B, Caby, S, Oger, F, Sasorith, S, Wurtz, JM and Pierce, RJ (2005) The monomeric orphan nuclear receptor Schistosoma mansoni Ftz-F1 dimerizes specifically and functionally with the schistosome RXR homologue, SmRXR1. Biochemical and Biophysical Research Communications 327, 10721082. https://doi.org/10.1016/j.bbrc.2004.12.101.CrossRefGoogle ScholarPubMed
Bertin, B, Oger, F, Cornette, J, Caby, S, Noël, C, Capron, M, Fantappie, MR, Rumjanek, FD and Pierce, RJ (2006) Schistosoma mansoni CBP/p300 has a conserved domain structure and interacts functionally with the nuclear receptor SmFtz-F1. Molecular and Biochemical Parasitology 146, 180191. https://doi.org/10.1016/j.molbiopara.2005.12.006.CrossRefGoogle Scholar
Bjarnadóttir, TK, Gloriam, DE, Hellstrand, SH, Kristiansson, H, Fredriksson, R and Schiöth, HB (2006) Comprehensive repertoire and phylogenetic analysis of the G protein-coupled receptors in human and mouse. Genomics 88, 263273. https://doi.org/10.1016/j.ygeno.2006.04.001.CrossRefGoogle Scholar
Blanton, RE and Licate, LS (1992) Developmental regulation of protein synthesis in schistosomes. Molecular and Biochemical Parasitology 51, 201208. https://doi.org/10.1016/0166-6851(92)90070-z.CrossRefGoogle ScholarPubMed
Bologna, Z, Teoh, J-P, Bayoumi, AS, Tang, Y, and Kim, I-M (2017) Biased G protein-coupled receptor signaling: new player in modulating physiology and pathology. Biomolecules & Therapeutics 25, 1225. https://doi.org/10.4062/biomolther.2016.165.CrossRefGoogle ScholarPubMed
Bostic, JR and Strand, M (1996) Molecular cloning of a Schistosoma mansoni protein expressed in the gynecophoral canal of male worms. Molecular and Biochemical Parasitology 79, 7989. https://doi.org/10.1016/0166-6851(96)02640-0.CrossRefGoogle ScholarPubMed
Buddenborg, SK, Kamel, B, Hanelt, B, Bu, L, Zhang, SM, Mkoji, GM, Loker, ES (2019) The in vivo transcriptome of Schistosoma mansoni in the prominent vector species Biomphalaria pfeifferi with supporting observations from Biomphalaria glabrata. PLOS Neglected Tropical Diseases 13(9): e0007013. https://doi.org/10.1371/journal.pntd.0007013CrossRefGoogle ScholarPubMed
Bueding, E, Liu, CL and Rogers, SH (1972) Inhibition by metrifonate and dichlorvos of cholinesterases in schistosomes. British Journal of Pharmacology 46, 480487. https://doi.org/10.1111/j.1476-5381.1972.tb08145.x.CrossRefGoogle ScholarPubMed
Bueding, E (1950) Carbohydrate metabolism of Schistosoma mansoni. Journal of General Physiology 33, 475495.CrossRefGoogle ScholarPubMed
Bulynko, YA and BW, O’Malle (2011) Nuclear receptor coactivators: Structural and functional biochemistry. Biochemistry 50, 313328. https://doi.org/10.1021/bi101762x.CrossRefGoogle ScholarPubMed
Buro, C, Burmeister, C, Quack, T and Grevelding, CG (2017) Identification and first characterization of SmEps8, a potential interaction partner of SmTK3 and SER transcribed in the gonads of Schistosoma mansoni. Experimental Parasitology 180, 5563. https://doi.org/10.1016/j.exppara.2016.12.002.CrossRefGoogle ScholarPubMed
Callau-Vázquez, D, Pless, SA and Lynagh, T (2018) Investigation of agonist recognition and channel properties in a flatworm glutamate-gated chloride channel. Biochemistry 57, 13601368. https://doi.org/10.1021/acs.biochem.7b01245.CrossRefGoogle Scholar
Camacho, M and Agnew, A (1995) Schistosoma: Rate of glucose import is altered by acetylcholine interaction with tegumental acetylcholine receptors and acetylcholinesterase. Experimental Parasitology 81, 584591. https://doi.org/10.1006/expr.1995.1152.CrossRefGoogle ScholarPubMed
Camacho, M, Tarrab-Hazdai, R, Espinoza, B, Arnon, R and Agnew, A (1994) The amount of acetylcholinesterase on the parasite surface reflects the differential sensitivity of schistosome species to metrifonate. Parasitology 108:153–60. https://doi.org/10.1017/s0031182000068244CrossRefGoogle ScholarPubMed
Carlo, JM, Osman, A, Niles, EG, Wu, W, Fantappie, MR, Oliveira, FM and LoVerde, PT (2007) Identification and characterization of an R-Smad ortholog (SmSmad1B) from Schistosoma mansoni. The FEBS Journal 274:4075–93. https://doi.org/10.1111/j.1742-4658.2007.05930.xCrossRefGoogle ScholarPubMed
Castanotto, D, Lingeman, R, Riggs, AD and Rossi, JJ (2009) CRM1 mediates nuclear-cytoplasmic shuttling of mature microRNAs. Proceedings of the National Academy of Sciences 106:21655–9. https://doi.org/10.1073/pnas.0912384106CrossRefGoogle ScholarPubMed
Cattaneo, F, Guerra, G, Parisi, M, De Marinis, M, Tafuri, D, Cinelli, M and Ammendola, R (2014) Cell-surface receptors transactivation mediated by G protein-coupled receptors. International Journal of Molecular Sciences 15, 1970019728. https://doi.org/10.3390/ijms151119700.CrossRefGoogle ScholarPubMed
Caveney, S, Cladman, W, Verellen, L and Donly, C (2006) Ancestry of neuronal monoamine transporters in the Metazoa. Journal of Experimental Biology. 209, 48584868. https://doi.org/10.1242/jeb.02607.CrossRefGoogle ScholarPubMed
Chan, JD, Cupit, PM, Gunaratne, GS, McCorvy, JD, Yang, Y, Stoltz, K, Webb, TR, Dosa, PI, Roth, BL, Abagyan, R, Cunningham, C and Marchant, JS (2017) The anthelmintic praziquantel is a human serotoninergic G-protein-coupled receptor ligand. Nature Communications 8, 1910. https://doi.org/10.1038/s41467-017-02084-0.CrossRefGoogle ScholarPubMed
Chen, L, Vasoya, RP, Toke, NH, Parthasarathy, A, Luo, S, Chiles, E, Flores, J, Gao, N, Bonder, EM, Su, X and Verzi, MP (2020) HNF4 regulates fatty acid oxidation and is required for renewal of intestinal stem cells in mice. Gastroenterology 158, 985999. e9. https://doi.org/10.1053/j.gastro.2019.11.031.CrossRefGoogle ScholarPubMed
Chen, R, Wang, J, Gradinaru, I, Vu, HS, Geboers, S, Naidoo, J, Ready, JM, Williams, NS, DeBerardinis, RJ, Ross, EM and Collins, JJ (2022) A male-derived nonribosomal peptide pheromone controls female schistosome development. Cell 185, 15061520. e17. https://doi.org/10.1016/j.cell.2022.03.017.CrossRefGoogle ScholarPubMed
Chen, PY, Qin, L and Simons, M (2023) TGFβ signaling pathways in human health and disease. Frontiers in Molecular Biosciences 10, 1113061. https://doi.org/10.3389/fmolb.2023.1113061.CrossRefGoogle ScholarPubMed
Cheng, G, Li, X, Qin, F, Xu, R, Zhang, Y, Liu, J, Gu, S and Jin, Y (2019) Functional analysis of theFrzb2 gene in Schistosoma japonicum. Veterinary Research 50, 108. https://doi.org/10.1186/s13567-019-0716-1.CrossRefGoogle ScholarPubMed
Childs, JE, Shirazian, D, Gloer, JB and Schiller, EL (1986) In vitro orientation of male Schistosoma mansoni to extracts derived from female schistosomes. Journal of Chemical Ecology 12, 17291738. https://doi.org/10.1007/BF01022378.CrossRefGoogle Scholar
Collins, JJ III, Wang, B, Lambrus, BG, Tharp, ME, Iyer, H and Newmark, PA (2013) Adult somatic stem cells in the human parasite Schistosoma mansoni. Nature 494, 476479. https://doi.org/10.1038/nature11924.CrossRefGoogle ScholarPubMed
Dale, HH (1914) The action of certain esters and ethers of choline, and their relation to muscarine. Journal of Pharmacology and Experimental Therapeutics 6, 147190.CrossRefGoogle Scholar
Davies, SJ, Shoemaker, CB and Pearce, EJ (1998) A divergent member of the transforming growth factor beta receptor family from Schistosoma mansoni is expressed on the parasite surface membrane. Journal of Biological Chemistry 273, 1123411240. https://doi.org/10.1074/jbc.273.18.11234.CrossRefGoogle ScholarPubMed
Day, TA, Chen, GZ, Miller, C, Tian, M, Bennett, JL and Pax, RA (1996) Cholinergic inhibition of muscle fibres isolated from Schistosoma mansoni (Trematoda: Digenea). Parasitology 113, 5561. https://doi.org/10.1017/s0031182000066270.CrossRefGoogle ScholarPubMed
de Jong-Brink, M, ter Maat, and Tensen, CP (2001) NPY in invertebrates: Molecular answers to altered functions during evolution. Peptides 22, 309315. https://doi.org/10.1016/s0196-9781(01)00332-1.CrossRefGoogle ScholarPubMed
De Mendonça, RL, Bouton, D, Bertin, B, Escriva, H, Noël, C, Vanacker, JM, Cornette, J, Laudet, V and Pierce, RJ (2002) A functionally conserved member of the FTZ-F1 nuclear receptor family from Schistosoma mansoni. European Journal of Biochemistry 269, 57005711. https://doi.org/10.1046/j.1432-1033.2002.03287.x.CrossRefGoogle ScholarPubMed
de Mendonça, RL, Escriva, H, Bouton, D, Zelus, D, Vanacker, JM, Bonnelye, E, Cornette, J, Pierce, RJ and Laudet, V (2000) Structural and functional divergence of a nuclear receptor of the RXR family from the trematode parasite Schistosoma mansoni. European Journal of Biochemistry 267, 32083219. https://doi.org/10.1046/j.1432-1327.2000.01344.x.CrossRefGoogle ScholarPubMed
Du, X, McManus, DP, Fogarty, CE, Jones, MK and You, H (2022) Schistosoma mansoni fibroblast growth factor receptor A orchestrates multiple functions in schistosome biology and in the host-parasite interplay. Frontiers in Immunology 13, 868077. https://doi.org/10.3389/fimmu.2022.868077.CrossRefGoogle ScholarPubMed
Du, X, McManus, DP, French, JD, Collinson, N, Sivakumaran, H, MacGregor, SR, Fogarty, CE, Jones, MK and You, H (2023) CRISPR interference for sequence-specific regulation of fibroblast growth factor receptor A in Schistosoma mansoni. Frontiers in Immunology 13:1105719. https://doi.org/10.3389/fimmu.2022.1105719CrossRefGoogle ScholarPubMed
Dufour, V, Beech, RN, Wever, C, Dent, JA and Geary, TG (2013) Molecular cloning and characterization of novel glutamate-gated chloride channel subunits from Schistosoma mansoni. PLOS Pathogens 9(8), e1003586. https://doi.org/10.1371/journal.ppat.1003586.CrossRefGoogle ScholarPubMed
Echeverria, PC and Picard, D (2010) Molecular chaperones, essential partners of steroid hormone receptors for activity and mobility. Biochimica et Biophysica Acta 1803, 641649. https://doi.org/10.1016/j.bbamcr.2009.11.012.CrossRefGoogle ScholarPubMed
El Naga, IF, Eissa, MM, Mossallam, SF and El-Halim, SI (2010) Inheritance of Schistosoma mansoni infection incompatibility in Biomphalaria alexandrina snails. Memórias do Instituto Oswaldo Cruz 105, 149154. https://doi.org/10.1590/s0074-02762010000200007.CrossRefGoogle ScholarPubMed
el Zawawy, LA, el Nassery, SF, al Azzouni, MZ, Abou el Naga, IF, el Temsahi, MM and Awadalla, HN (1995) A study on patients with eosinophilia of suspected parasitic origin. Journal of the Egyptian Society of Parasitology 25, 245255.Google Scholar
Elhenawy, AA, Ashour, RH, Nabih, N, Shalaby, NM, El-Karef, AA and Abou-El-Wafa, HS (2017) Insulin growth factor inhibitor as a potential new anti-schistosoma drug: An in vivo experimental study. Biomedicine and Pharmacotherapy 95, 13461358. https://doi.org/10.1016/j.biopha.2017.09.015.CrossRefGoogle ScholarPubMed
El-Sakkary, N, Chen, S, Arkin, MR, Caffrey, CR and Ribeiro, P (2018) Octopamine signaling in the metazoan pathogen Schistosoma mansoni: Localization, small-molecule screening and opportunities for drug development. Disease Models & Mechanisms 11(7), dmm033563. https://doi.org/10.1242/dmm.033563.CrossRefGoogle ScholarPubMed
El-Shabasy, EA, Saleh, MA, Said, AE and Reda, ES (2024) Evaluation of Schistosoma mansoni nervous system using confocal laser electron microscopy: Nerve sensilla and FMRFamide while referring to F-actin abundance. Egyptian Journal of Basic and Applied Sciences 11, 281296. https://doi.org/10.1080/2314808X.2024.2335853.CrossRefGoogle Scholar
El-Shehabi, F, Vermeire, J, Timothy, P and Yoshino, TPR (2009) Developmental expression analysis and immunolocalization of a biogenic amine receptor in Schistosoma mansoni. Experimental Parasitology 122, 1727. https://doi.org/10.1016/j.exppara.2009.01.001.CrossRefGoogle ScholarPubMed
El-Shehabi, F and Ribeiro, P (2010) Histamine signalling in Schistosoma mansoni: Immunolocalisation and characterisation of a new histamine-responsive receptor (SmGPR-2). International Journal for Parasitology 40, 1395–406. https://doi.org/10.1016/j.ijpara.2010.04.006CrossRefGoogle ScholarPubMed
El-Shehabi, F, Taman, A, Moali, LS, El-Sakkary, N and Ribeiro, P (2012) A novel G protein-coupled receptor of Schistosoma mansoni (smGPR-3) is activated by dopamine and is widely expressed in the nervous system. PLOS Neglected Tropical Diseases 6, e1523. https://doi.org/10.1371/journal.pntd.0001523.CrossRefGoogle Scholar
Escobedo, G, Roberts, CW, Carrero, JC and Morales-Montor, J (2005) Parasite regulation by host hormones: An old mechanism of host exploitation? Trends in Parasitology 21,588593. https://doi.org/10.1016/j.pt.2005.09.013.CrossRefGoogle ScholarPubMed
Espinoza, B, Tarrab-Hazdai, R, Silman, I, and Arnon, R (1988) Acetylcholinesterase in Schistosoma mansoni is anchored to the membrane via covalently attached phosphatidylinositol. Molecular and Biochemical Parasitology 29, 171179. https://doi.org/10.1016/0166-6851(88)90072-2.CrossRefGoogle Scholar
Espinoza, B, Silman, I, Arnon, R and Tarrab-Hazdai, R (1991) Phosphatidylinositol-specific phospholipase C induces biosynthesis of acetylcholinesterase via diacylglycerol in Schistosoma mansoni. European Journal of Biochemistry 195, 863870. https://doi.org/10.1111/j.1432-1033.1991.tb15776.x.CrossRefGoogle ScholarPubMed
Fantappie, MR, Freebern, WJ, Osman, A, LaDuca, J, Niles, EG, and LoVerde, PT (2001) Evaluation of Schistosoma mansoni retinoid X receptor (SmRXR1 and SmRXR2) activity and tissue distribution. Molecular and Biochemical Parasitology 115, 8799. https://doi.org/10.1016/s0166-6851(01)00274-2.CrossRefGoogle ScholarPubMed
Fantappié, MR, Bastos de Oliveira, FM, de Moraes Maciel, R, Rumjanek, FD, Wu, W and LoVerde, PT (2008a) Cloning of SmNCoA-62, a novel nuclear receptor co-activator from Schistosoma mansoni: Assembly of a complex with a SmRXR1/SmNR1 heterodimer, SmGCN5 and SmCBP1. International Journal for Parasitology 38, 11331347. https://doi.org/10.1016/j.ijpara.2008.02.003.CrossRefGoogle ScholarPubMed
Fantappié, MR, Furtado, DR, Rumjanek, FD and LoVerde, PT (2008b) A unique nuclear receptor direct repeat 17 (DR17) is present within the upstream region of Schistosoma mansoni female-specific p14 gene. Biochemical and Biophysical Research Communications 371, 689693. https://doi.org/10.1016/j.bbrc.2008.04.125CrossRefGoogle ScholarPubMed
Figliuolo da Paz, VR, Figueiredo-Vanzan, D and Dos Santos Pyrrho, A (2019) Interaction and involvement of cellular adhesion molecules in the pathogenesis of Schistosoma mansoni. Immunology Letters 206, 1118. https://doi.org/10.1016/j.imlet.2018.11.011.CrossRefGoogle Scholar
Forrester, SG, Warfel, PW and Pearce, EJ (2004) Tegumental expression of a novel type II receptor serine/threonine kinase (SmRK2) in Schistosoma mansoni. Molecular and Biochemical Parasitology 136, 149156. https://doi.org/10.1016/j.molbiopara.2004.03.007.CrossRefGoogle ScholarPubMed
Fredriksson, R, Lagerström, MC, Lundin, LG and Schiöth, HB (2003) The G-protein-coupled receptors in the human genome form five main families: phylogenetic analysis, paralogon groups, and fingerprints. Molecular Pharmacology 63, 12561272. https://doi.org/10.1124/mol.63.6.1256.CrossRefGoogle ScholarPubMed
Freebern, WJ, Niles, EG and LoVerde, PT (1999a) RXR-2, a member of the retinoid x receptor family in Schistosoma mansoni. Gene 233, 3338. https://doi.org/10.1016/s0378-1119(99)00161-4.CrossRefGoogle ScholarPubMed
Freebern, WJ, Osman, A, Niles, EG, Christen, L and LoVerde, PT (1999b) Identification of a cDNA encoding a retinoid X receptor homologue from Schistosoma mansoni: Evidence for a role in female-specific gene expression. Journal of Biological Chemistry 274, 45774585. https://doi.org/10.1074/jbc.274.8.4577.CrossRefGoogle ScholarPubMed
Freitas, TC, Jung, E and Pearce, EJ (2007) TGF-beta signaling controls embryo development in the parasitic flatworm Schistosoma mansoni. PLOS Pathogens 3(4), e52. https://doi.org/10.1371/journal.ppat.0030052.CrossRefGoogle ScholarPubMed
Freitas, TC, Jung, E and Pearce, EJ (2009) A bone morphogenetic protein homologue in the parasitic flatworm, Schistosoma mansoni. International Journal for Parasitology 39, 281287. https://doi.org/10.1016/j.ijpara.2008.08.001.CrossRefGoogle ScholarPubMed
Frooninckx, L, Van Rompay, L, Temmerman, L, Van Sinay, E, Beets, I, Janssen, T, Husson, SJ and Schoofs, L (2012) Neuropeptide GPCRs in C. elegans. Frontiers in Endocrinology 3, 167. https://doi.org/10.3389/fendo.2012.00167.CrossRefGoogle Scholar
Fu, G, Wang, W and Luo, BH (2012) Overview: Structural biology of integrins. Methods in Molecular Biology 757, 8199. https://doi.org/10.1007/978-1-61779-166-6_7.CrossRefGoogle ScholarPubMed
García-Tobilla, P, Solórzano, SR, Salido-Guadarrama, I, González-Covarrubias, V, Morales-Montor, G, Díaz-Otañez, CE and Rodríguez-Dorantes, M (2016) SFRP1 repression in prostate cancer is triggered by two different epigenetic mechanisms. Gene 593, 292301. https://doi.org/10.1016/j.gene.2016.08.030.CrossRefGoogle ScholarPubMed
Gelmedin, V, Morel, M, Hahnel, S, Cailliau, K, Dissous, C and Grevelding, CG (2017) Evidence for integrin-Venus kinase receptor 1 alliance in the ovary of Schistosoma mansoni females controlling cell survival. PLOS Pathogens 13(1), e1006147. https://doi.org/10.1371/journal.ppat.1006147.CrossRefGoogle ScholarPubMed
Giannini, AL, Caride, EC, Braga, VM and Rumjanek, FD (1995) F-10 nuclear binding proteins of Schistosoma mansoni: Structural and functional features. Parasitology 110, 155161. https://doi.org/10.1017/s0031182000063915.CrossRefGoogle ScholarPubMed
Gouignard, N, Vanderstraete, M, Cailliau, K, Lescuyer, A, Browaeys, E and Dissous, C (2012) Schistosoma mansoni: Structural and biochemical characterization of two distinct Venus kinase receptors. Experimental Parasitology 132, 3239. https://doi.org/10.1016/j.exppara.2011.05.007.CrossRefGoogle ScholarPubMed
Grapa, CM, Mocan, T, Gonciar, D, Zdrehus, C, Mosteanu, O, Pop, T and Mocan, L (2019) Epidermal growth factor receptor and its role in pancreatic cancer treatment mediated by nanoparticles. International Journal of Nanomedicine 14, 96939706. https://doi.org/10.2147/ijn.S226628.CrossRefGoogle ScholarPubMed
Gurevich, EV, Gainetdinov, RR and Gurevich, VV (2016) G protein-coupled receptor kinases as regulators of dopamine receptor functions. Pharmacological Research 111, 116. https://doi.org/10.1016/j.phrs.2016.05.010.CrossRefGoogle ScholarPubMed
Hahnel, S, Quack, T, Parker-Manuel, SJ, Lu, Z, Vanderstraete, M, Morel, M, Dissous, C, Cailliau, K and Grevelding, CG (2014) Gonad RNA-specific qRT-PCR analyses identify genes with potential functions in schistosome reproduction such as SmFz1 and SmFGFRs. Frontiers in Genetics 5, 170. https://doi.org/10.3389/fgene.2014.00170.CrossRefGoogle ScholarPubMed
Hahnel, S, Wheeler, N, Lu, Z, Wangwiwatsin, A, McVeigh, P, Maule, A, Berriman, M, Day, T, Ribeiro, P and Grevelding, CG (2018) Tissue-specific transcriptome analyses provide new insights into GPCR signalling in adult Schistosoma mansoni. PLOS Pathogens 14(1), e1006718. https://doi.org/10.1371/journal.ppat.1006718.CrossRefGoogle ScholarPubMed
Halton, DW and Maule, AG (2004) Flatworm nerve-muscle: Structural and functional analysis. Canadian Journal of Zoology 82, 316333. https://doi.org/10.1139/z03-221.CrossRefGoogle Scholar
Hamdan, FF and Ribeiro, P (1999) Characterization of a stable form of tryptophan hydroxylase from the human parasite Schistosoma mansoni. Journal of Biological Chemistry 274, 2174621754. https://doi.org/10.1074/jbc.274.31.21746.CrossRefGoogle ScholarPubMed
Hamdan, FF, Abramovitz, M, Mousa, A, Xie, J, Durocher, Y and Ribeiro, P (2002) A novel Schistosoma mansoni G protein-coupled receptor is responsive to histamine. Molecular and Biochemical Parasitology 119, 7586. https://doi.org/10.1016/s0166-6851(01)00400-5.CrossRefGoogle ScholarPubMed
Hammouda, NA, Abou el Naga, IF, el Temsahi, MM and Sharaf, IA (1994) Schistosoma mansoni: A comparative study on two cercarial transformation methods. Journal of the Egyptian Society of Parasitology 24, 479486.Google ScholarPubMed
Harburger, DS and Calderwood, DA (2009) Integrin signalling at a glance. Journal of Cell Science 122, 159163. https://doi.org/10.1242/jcs.018093.CrossRefGoogle ScholarPubMed
Harder, A (2002) Chemotherapeutic approaches to schistosomes: Current knowledge and outlook. Parasitology Research 88, 395397. https://doi.org/10.1007/s00436-001-0588-x.CrossRefGoogle ScholarPubMed
Hata, A and Chen, YG (2016) TGF-β signaling from receptors to Smads. Cold Spring Harbor Perspectives in Biology 8(9): a022061. https://doi.org/10.1101/cshperspect.a022061.CrossRefGoogle ScholarPubMed
Hering, H and Sheng, M (2002) Direct interaction of Frizzled-1, -2, -4, and -7 with PDZ domains of PSD-95. FEBS Letters 521, 185189. https://doi.org/10.1016/s0014-5793(02)02831-4.CrossRefGoogle ScholarPubMed
Hill, CA, Sharan, S and Watts, VJ (2018) Genomics, GPCRs and new targets for the control of insect pests and vectors. Current Opinion in Insect Science 30, 99106. https://doi.org/10.1016/j.cois.2018.08.010.CrossRefGoogle ScholarPubMed
Hoffmann, KF, Davis, EM, Fischer, ER and Wynn, TA (2001) The guanine protein coupled receptor rhodopsin is developmentally regulated in the freeliving stages of Schistosoma mansoni. Molecular and Biochemical Parasitology 112, 113123. https://doi.org/10.1016/s0166-6851(00)00352-2.CrossRefGoogle ScholarPubMed
Hofmann, L and Palczewski, K (2015) The G protein-coupled receptor rhodopsin: A historical perspective. Methods in Molecular Biology 1271, 318. https://doi.org/10.1007/978-1-4939-2330-4_1.CrossRefGoogle Scholar
Hong, F, Pan, S, Guo, Y, Xu, P and Zhai, Y (2019) PPARs as nuclear receptors for nutrient and energy metabolism. Molecules 24:2545. https://doi.org/10.3390/molecules24142545CrossRefGoogle ScholarPubMed
Hu, R, Wu, W, Niles, EG and LoVerde, PT (2006a) Isolation and characterization of Schistosoma mansoni constitutive androstane receptor. Molecular and Biochemical Parasitology 148, 3143. https://doi.org/10.1016/j.molbiopara.2006.02.017.CrossRefGoogle ScholarPubMed
Hu, R, Niles, EG and LoVerde, PT (2006b) DNA binding and transactivation properties of the Schistosoma mansoni constitutive androstane receptor homologue. Molecular and Biochemical Parasitology 150, 174185. https://doi.org/10.1016/j.molbiopara.2006.07.011.CrossRefGoogle ScholarPubMed
Hu, R, Wu, W, Niles, EG and LoVerde, PT (2006c) SmTR2/4, a Schistosoma mansoni homologue of TR2/TR4 orphan nuclear receptor. International Journal for Parasitology 36, 11131122. https://doi.org/10.1016/j.ijpara.2006.06.003.CrossRefGoogle ScholarPubMed
Hynes, RO (2002) Integrins: Bidirectional, allosteric signaling machines. Cell 110, 673687. https://doi.org/10.1016/s0092-8674(02)00971-6.CrossRefGoogle ScholarPubMed
Jones, AK, Bentley, GN, Oliveros Parra, WG and Agnew, A (2002) Molecular characterization of an acetylcholinesterase implicated in the regulation of glucose scavenging by the parasite Schistosoma. The FASEB Journal 16, 441443. https://doi.org/10.1096/fj.01-0683fje.CrossRefGoogle ScholarPubMed
Kamara, IK, Thao, JT, Kaur, K, Wheeler, NJ and Chan, JD (2023) Annotation of G-protein coupled receptors in the genomes of parasitic blood flukes. microPublication Biology 2023, 10.17912/micropub.biology.000704. https://doi.org/10.17912/micropub.biology.000704.Google ScholarPubMed
Kapp, K, Knobloch, J, Schüssler, P, Sroka, S, Lammers, R, Kunz, W and Grevelding, CG (2004) The Schistosoma mansoni Src kinase TK3 is expressed in the gonads and likely involved in cytoskeletal organization. Molecular and Biochemical Parasitology 138, 171182. https://doi.org/10.1016/j.molbiopara.2004.07.010.CrossRefGoogle ScholarPubMed
Kaur, S, Jobling, S, Jones, CS, Noble, LR, Routledge, EJ and Lockyer, AE (2015) The nuclear receptors of Biomphalaria glabrata and Lottia gigantea: implications for developing new model organisms. PLOS One 10(4), e0121259. https://doi.org/10.1371/journal.pone.0121259.CrossRefGoogle ScholarPubMed
Keramidas, A, Moorhouse, AJ, Pierce, KD, Schofield, PR and Barry, PH (2002) Cation-selective mutations in the M2 domain of the inhibitory glycine receptor channel reveal determinants of ion-charge selectivity. Journal of General Physiology 119, 393410. https://doi.org/10.1085/jgp.20028552.CrossRefGoogle ScholarPubMed
Khayath, N, Vicogne, J, Ahier, A, BenYounes, A, Konrad, C, Trolet, J, Viscogliosi, E, Brehm, K and Dissous, C (2007) Diversification of the insulin receptor family in the helminth parasite Schistosoma mansoni. The FEBS Journal 274, 659676. https://doi.org/10.1111/j.1742-4658.2006.05610.x.CrossRefGoogle ScholarPubMed
Kim, JJ, and Accili, D (2002) Signalling through IGF-I and insulin receptors: Where is the specificity. Growth Hormone & IGF Research 12, 8490. https://doi.org/10.1054/ghir.2002.0265.CrossRefGoogle ScholarPubMed
Knobloch, J, Rossi, A, Osman, A, LoVerde, PT, Klinkert, MQ and Grevelding, CG (2004) Cytological and biochemical evidence for a gonad-preferential interplay of SmFKBP12 and SmTbetaR-I in Schistosoma mansoni. Molecular and Biochemical Parasitology 138, 227236. https://doi.org/10.1016/j.molbiopara.2004.09.006.CrossRefGoogle ScholarPubMed
Knobloch, J, Beckmann, S, Burmeister, C, Quack, T and Grevelding, CG (2007) Tyrosine kinase and cooperative TGF beta signaling in the reproductive organs of Schistosoma mansoni. Experimental Parasitology 117, 318336. https://doi.org/10.1016/j.exppara.2007.04.006.CrossRefGoogle ScholarPubMed
Köhler, A and Hurt, E (2007) Exporting RNA from the nucleus to the cytoplasm. Nature Reviews Molecular Cell Biology 8, 761773. https://doi.org/10.1038/nrm2255.CrossRefGoogle ScholarPubMed
Kojetin, DJ, Matta-Camacho, E, Hughes, TS, Srinivasan, S, Nwachukwu, JC, Cavett, V, Nowak, J, Chalmers, MJ, Marciano, DP, Kamenecka, TM, Shulman, AI, Rance, M, Griffin, PR, Bruning, JB and Nettles, KW (2015) Structural mechanism for signal transduction in RXR nuclear receptor heterodimers. Nature Communications 6, 8013. https://doi.org/10.1038/ncomms9013.CrossRefGoogle ScholarPubMed
Kreshchenko, ND (2008) Functions of flatworm neuropeptides NPF, GYIRF and FMRF in course of pharyngeal regeneration of anterior body fragments of planarian, Girardia tigrina. Acta Biologica Hungarica 59, 199207. https://doi.org/10.1556/ABiol.59.2008.suppl.29.CrossRefGoogle ScholarPubMed
Kristiansen, K (2004) Molecular mechanisms of ligand binding, signaling, and regulation within the superfamily of G-protein-coupled receptors: Molecular modeling and mutagenesis approaches to receptor structure and function. Pharmacology & Therapeutics 103, 2180. https://doi.org/10.1016/j.pharmthera.2004.05.002.CrossRefGoogle ScholarPubMed
Kryszkowski, W and Boczek, T (2021) The G protein-coupled glutamate receptors as novel molecular targets in schizophrenia treatment - a narrative review. Journal of Clinical Medicine 10, 1475. https://doi.org/10.3390/jcm10071475.CrossRefGoogle Scholar
Kunz, W (2001) Schistosome male-female interaction: Induction of germ-cell differentiation. Trends in Parasitology 17, 227231. https://doi.org/10.1016/s1471-4922(01)01893-1.CrossRefGoogle ScholarPubMed
Lackner, DH and Bähler, J (2008) Translational control of gene expression from transcripts to transcriptomes. International Review of Cell and Molecular Biology 271, 199251. https://doi.org/10.1016/S1937-6448(08)01205-7.CrossRefGoogle ScholarPubMed
Lagerström, MC and Schiöth, HB (2008) Structural diversity of G protein-coupled receptors and significance for drug discovery. Nature Reviews Drug Discovery 7, 339357. https://doi.org/10.1038/nrd2518.CrossRefGoogle ScholarPubMed
Lazar, MA (2017) Maturing of the nuclear receptor family. Journal of Clinical Investigations 127, 11231125. https://doi.org/10.1172/JCI92949.CrossRefGoogle ScholarPubMed
Lee, J, Chong, T and Newmark, PA (2020) The esophageal gland mediates host immune evasion by the human parasite Schistosoma mansoni. Proceedings of the National Academy of Sciences 117, 1929919309. https://doi.org/10.1073/pnas.2006553117.CrossRefGoogle ScholarPubMed
Li, X, Weth, O, Haimann, M, Möscheid, MF, Huber, TS and Grevelding, CG (2024) Rhodopsin orphan GPCR20 interacts with neuropeptides and directs growth, sexual differentiation, and egg production in female Schistosoma mansoni. Microbiology Spectrum 12, 12:e0219323. https://doi.org/10.1128/spectrum.02193-23.Google ScholarPubMed
LoVerde, PT, Niles, EG, Osman, A and Wu, WJ (2004) Schistosoma mansoni male-female interactions. Canadian Journal of Zoology 82, 357374. https://cdnsciencepub.com/doi/10.1139/z03-217.CrossRefGoogle Scholar
LoVerde, PT, Osman, A and Hinck, A (2007) Schistosoma mansoni: TGF-beta signaling pathways. Experimental Parasitology 117, 304317. https://doi.org/10.1016/j.exppara.2007.06.002.CrossRefGoogle ScholarPubMed
Lu, C, Wu, W, Niles, EG and LoVerde, PT (2006) Identification and characterization of a novel fushi tarazu factor 1 (FTZ-F1) nuclear receptor in Schistosoma mansoni. Molecular and Biochemical Parasitology 150, 2536. https://doi.org/10.1016/j.molbiopara.2006.06.005.CrossRefGoogle ScholarPubMed
Lu, Z, Sessler, F, Holroyd, N, Hahnel, S, Quack, T, Berriman, M and Grevelding, CG (2016) Schistosome sex matters: A deep view into gonad-specific and pairing-dependent transcriptomes reveals a complex gender interplay. Scientific Reports 6, 31150. https://doi.org/10.1038/srep31150.CrossRefGoogle ScholarPubMed
Luttrell, LM (2008) Reviews in molecular biology and biotechnology: Transmembrane signaling by G protein-coupled receptors. Molecular Biotechnology 39, 239264. https://doi.org/10.1007/s12033-008-9031-1.CrossRefGoogle ScholarPubMed
MacDonald, K, Buxton, S, Kimber, MJ, Day, TA, Robertson, AP and Ribeiro, P (2014) Functional characterization of a novel family of acetylcholine-gated chloride channels in Schistosoma mansoni. PLOS Pathogens 10(6), e1004181. https://doi.org/10.1371/journal.ppat.1004181.CrossRefGoogle ScholarPubMed
MacDonald, K, Kimber, MJ, Day, TA and Ribeiro, P (2015) A constitutively active G protein-coupled acetylcholine receptor regulates motility of larval Schistosoma mansoni. Molecular and Biochemical Parasitology 202, 2937. https://doi.org/10.1016/j.molbiopara.2015.09.001.CrossRefGoogle ScholarPubMed
Mackmull, MT, Klaus, B, Heinze, I, Chokkalingam, M, Beyer, A, Russell, RB, Ori, A and Beck, M (2017) Landscape of nuclear transport receptor cargo specificity. Molecular Systems Biology 13, 962. https://doi.org/10.15252/msb.20177608.CrossRefGoogle ScholarPubMed
Maharjan, S, Kirk, RS, Lawton, SP and Walker, AJ (2023) Human growth factor mediated signalling through lipid rafts regulates stem cell proliferation, development and survival of Schistosoma mansoni. Open Biology 13, 230262. https://doi.org/10.1098/rsob.230262.Google Scholar
Mathavan, I, Liu, LJ, Robinson, SW, El-Sakkary, N, Elatico, AJJ, Gomez, D, Nellas, R, Owens, RJ, Zuercher, W, Navratilova, I, Caffrey, CR and Beis, K (2022) Identification of inhibitors of the Schistosoma mansoni VKR2 kinase domain. ACS Medicinal Chemistry Letters 13, 17151722. https://doi.org/10.1021/acsmedchemlett.2c00248.CrossRefGoogle ScholarPubMed
McGonigle, S, Beall, MJ, Feeney, EL and Pearce, EJ (2001) Conserved role for 14-3-3epsilon downstream of type I TGFbeta receptors. FEBS Letters 490, 6569. https://doi.org/10.1016/s0014-5793(01)02133-0.CrossRefGoogle ScholarPubMed
McGonigle, S, Beall, MJ and Pearce, EJ (2002) Eukaryotic initiation factor 2 alpha subunit associates with TGF beta receptors and 14-3-3 epsilon and acts as a modulator of the TGF beta response. Biochemistry 41, 579587. https://doi.org/10.1021/bi011407z.CrossRefGoogle ScholarPubMed
McKenzie, M, Kirk, RS and Walker, AJ (2018) Glucose uptake in the human pathogen Schistosoma mansoni is regulated through Akt/protein kinase B signaling. The Journal of Infectious Diseases 218, 152164. https://doi.org/10.1093/infdis/jix654.CrossRefGoogle ScholarPubMed
McVeigh, P, Mair, GR, Atkinson, L, Ladurner, P, Zamanian, M, Novozhilova, E, Marks, NJ, Day, TA and Maule, AG (2009) Discovery of multiple neuropeptide families in the phylum Platyhelminthes. International Journal for Parasitology 39, 12431252. https://doi.org/10.1016/j.ijpara.2009.03.005.CrossRefGoogle ScholarPubMed
Mendonça-Silva, DL, Pessôa, RF and Noël, F (2002) Evidence for the presence of glutamatergic receptors in adult Schistosoma mansoni. Biochemical Pharmacology 64, 13371344. https://doi.org/10.1016/s0006-2952(02)01358-8.CrossRefGoogle ScholarPubMed
Miao, Y and McCammon, JA (2016) G-protein coupled receptors: advances in simulation and drug discovery. Current Opinion in Structural Biology 41, 8389. https://doi.org/10.1016/j.sbi.2016.06.008.CrossRefGoogle Scholar
Mingot, JM, Bohnsack, MT, Jäkle, U and Görlich, D (2004) Exportin 7 defines a novel general nuclear export pathway. The EMBO Journal 23, 32273236. https://doi.org/10.1038/sj.emboj.7600338.CrossRefGoogle ScholarPubMed
Moescheid, MF, Lu, Z, Soria, CD, Quack, T, Puckelwaldt, O, Holroyd, N, Holzaepfel, P, Haeberlein, S, Rinaldi, G, Berriman, M and Grevelding, CG (2024) The retinoic acid family-like nuclear receptor SmRAR identified by single-cell transcriptomics of ovarian cells controls oocyte differentiation in Schistosoma mansoni. Nucleic Acids Research 16, gkae1228. https://doi.org/10.1093/nar/gkae1228.CrossRefGoogle Scholar
Mogahed, NMFH, El-Temsahy, MM, Abou-El-Naga, IF, Makled, S, Sheta, E and Ibrahim, EI (2023) Loading praziquantel within solid lipid nanoparticles improved its schistosomicidal efficacy against the juvenile stage. Experimental Parasitology 251, 108552. https://doi.org/10.1016/j.exppara.2023.108552.CrossRefGoogle ScholarPubMed
Morel, M, Vanderstraete, M, Cailliau, K, Lescuyer, A, Lancelot, J and Dissous, C (2014) Compound library screening identified Akt/PKB kinase pathway inhibitors as potential key molecules for the development of new chemotherapeutics against schistosomiasis. International Journal for Parasitology Drugs Drug Resistance 4, 256266. https://doi.org/10.1016/j.ijpddr.2014.09.004.CrossRefGoogle ScholarPubMed
Morrison, DD, Vande Waa, EA and Bennett, JL (1986) Effects of steroids and steroid synthesis inhibitors on fecundity of Schistosoma mansoni in vitro. Journal of Chemical Ecology 12, 19011908. https://doi.org/10.1007/BF01022391.CrossRefGoogle Scholar
Murphy, D, Dancis, B and Brown, JR (2008) The evolution of core proteins involved in microRNA biogenesis. BMC Evolutionary Biology 8, 92. https://doi.org/10.1186/1471-2148-8-92.CrossRefGoogle ScholarPubMed
Nirde, P, Torpier, G, De Reggi, ML and Capron, A (1983) Ecdysone and 20 hydroxyecdysone: new hormones for the human parasite Schistosoma mansoni. FEBS Letters 151, 223227. https://doi.org/10.1016/0014-5793(83)80074-x.CrossRefGoogle ScholarPubMed
Niswender, CM and Conn, PJ (2010) Metabotropic glutamate receptors: Physiology, pharmacology, and disease. Annual Review of Pharmacology and Toxicology 50, 295322. https://doi.org/10.1146/annurev.pharmtox.011008.145533.CrossRefGoogle ScholarPubMed
Nordström, KJ, Lagerström, MC, Wallér, LM, Fredriksson, R and Schiöth, HB (2009) The Secretin GPCRs descended from the family of Adhesion GPCRs. Molecular Biology and Evolution 26, 7184. doi: 10.1093/molbev/msn228.CrossRefGoogle ScholarPubMed
Novac, N and Heinzel, T (2004) Nuclear receptors: Overview and classification. Current Drug Targets - Inflammation and Allergy 3, 335346. https://doi.org/10.2174/1568010042634541.Google Scholar
Nuclear Receptors Nomenclature Committee (1999) A unified nomenclature system for the nuclear receptor superfamily. Cell 97, 161163. https://doi.org/10.1016/s0092-8674(00)80726-6.CrossRefGoogle Scholar
Nusse, R (2015) Cell signalling: Disarming Wnt. Nature 519, 163164. https://doi.org/10.1038/nature14208.CrossRefGoogle ScholarPubMed
Okada, C, Yamashita, E, Lee, SJ, Shibata, S, Katahira, J, Nakagawa, A, Yoneda, Y and Tsukihara, T (2009) A high-resolution structure of the pre-microRNA nuclear export machinery. Science 326, 12751279. https://doi.org/10.1126/science.1178705.CrossRefGoogle ScholarPubMed
Oliveira, KC, Carvalho, ML, Verjovski-Almeida, S and LoVerde, PT (2012) Effect of human TGF-β on the gene expression profile of Schistosoma mansoni adult worms. Molecular and Biochemical Parasitology 183, 132139. https://doi.org/10.1016/j.molbiopara.2012.02.008.CrossRefGoogle ScholarPubMed
Ornitz, DM and Itoh, N (2015) The fibroblast growth factor signaling pathway. Wiley Interdisciplinary Reviews: Developmental Biology 4, 215266. https://doi.org/10.1002/wdev.176.CrossRefGoogle ScholarPubMed
Osman, A, Niles, EG, and LoVerde, PT (2001) Identification and characterization of a Smad2 homologue from Schistosoma mansoni, a transforming growth factor-beta signal transducer. Journal of Biological Chemistry 276, 1007210082. https://doi.org/10.1074/jbc.M005933200.CrossRefGoogle ScholarPubMed
Osman, A, Niles, EG and LoVerde, PT (2004) Expression of functional Schistosoma mansoni Smad4: Role in Erk-mediated transforming growth factor beta (TGF-beta) down-regulation. Journal of Biological Chemistry 279, 64746486. https://doi.org/10.1074/jbc.M310949200.CrossRefGoogle ScholarPubMed
Osman, A, Niles, EG, Verjovski-Almeida, S and LoVerde, PT (2006) Schistosoma mansoni TGF-beta receptor II: Role in host ligand-induced regulation of a schistosome target gene. PLOS Pathogens PLoS Pathogen 2(6), e54. https://doi.org/10.1371/journal.ppat.0020054.CrossRefGoogle ScholarPubMed
Patel, SR and Skafar, DF (2015) Modulation of nuclear receptor activity by the F domain. Molecular and Cellular Endocrinology 418, 298305. https://doi.org/10.1016/j.mce.2015.07.009.CrossRefGoogle ScholarPubMed
Patocka, N and Ribeiro, P (2007) Characterization of a serotonin transporter in the parasitic flatworm, Schistosoma mansoni: Cloning, expression and functional analysis. Molecular and Biochemical Parasitology 154, 125133. https://doi.org/10.1016/j.molbiopara.2007.03.010.CrossRefGoogle ScholarPubMed
Patocka, N, Sharma, N, Rashid, M and Ribeiro, P (2014) Serotonin signaling in Schistosoma mansoni: A serotonin-activated G protein-coupled receptor controls parasite movement. PLOS Pathogens 10(1), e1003878. https://doi.org/10.1371/journal.ppat.1003878.CrossRefGoogle ScholarPubMed
Pawlak, M, Lefebvre, P and Staels, B (2012) General molecular biology and architecture of nuclear receptors. Current Topics in Medical Chemistry 12, 486504. https://doi.org/10.2174/156802612799436641.CrossRefGoogle ScholarPubMed
Phan, P, Liang, D, Zhao, M, Wyeth, RC, Fogarty, C, Duke, MG, McManus, DP, Wang, T and Cummins, SF (2022) Analysis of rhodopsin G protein-coupled receptor orthologs reveals semiochemical peptides for parasite (Schistosoma mansoni) and host (Biomphalaria glabrata) interplay. Scientific Reports 12, 8243. https://doi.org/10.1038/s41598-022-11996-x.CrossRefGoogle ScholarPubMed
Popiel, I, Cioli, D and Erasmus, DA (1984) The morphology and reproductive status of female Schistosoma mansoni following separation from male worms. International Journal for Parasitology 14:183–90. https://doi.org/10.1016/0020-7519(84)90047-xCrossRefGoogle ScholarPubMed
Protasio, AV, Tsai, IJ, Babbage, A, Nichol, S, Hunt, M, Aslett, MA, De Silva, N, Velarde, GS, Anderson, TJ, Clark, RC, Davidson, C, Dillon, GP, Holroyd, NE, LoVerde, PT, Lloyd, C, McQuillan, J, Oliveira, G, Otto, TD, Parker-Manuel, SJ, Quail, MA, Wilson, RA, Zerlotini, A, Dunne, DW and Berriman, M (2012) A systematically improved high quality genome and transcriptome of the human blood fluke Schistosoma mansoni. PLOS Neglected Tropical Diseases 6(1), e1455. https://doi.org/10.1371/journal.pntd.0001455.CrossRefGoogle ScholarPubMed
Pyne, NJ and Pyne, S (2011) Receptor tyrosine kinase-G-protein-coupled receptor signalling platforms: Out of the shadow? Trends in Pharmacological Sciences 32, 443450. https://doi.org/10.1016/j.tips.2011.04.002.CrossRefGoogle ScholarPubMed
Ramachandran, H, Skelly, PJ and Shoemaker, CB (1996) The Schistosoma mansoni epidermal growth factor receptor homologue, SER, has tyrosine kinase activity and is localized in adult muscle. Molecular and Biochemical Parasitology 83, 110. https://doi.org/10.1016/s0166-6851(96)02731-4.CrossRefGoogle ScholarPubMed
Reiner, A and Levitz, J (2018) Glutamatergic signaling in the central nervous system: Ionotropic and metabotropic receptors in concert. Neuron 98, 10801098. https://doi.org/10.1016/j.neuron.2018.05.018.CrossRefGoogle ScholarPubMed
Ressurreição, M, Elbeyioglu, F, Kirk, RS, Rollinson, D, Emery, AM, Page, NM and Walker, AJ (2016) Molecular characterization of host-parasite cell signalling in Schistosoma mansoni during early development. Scientific Reports 6, 35614. https://doi.org/10.1038/srep35614.CrossRefGoogle ScholarPubMed
Ribeiro, P and Geary, T (2010) Neuronal signaling in schistosomes: Current status and prospects for postgenomics. Canadian Journal of Zoology 88, 122. https://doi.org/10.1139/Z09-126.CrossRefGoogle Scholar
Ribeiro, P and Patocka, N (2013) Neurotransmitter transporters in schistosomes: Structure, function and prospects for drug discovery. Parasitology International 62, 629638. https://doi.org/10.1016/j.parint.2013.06.003.CrossRefGoogle ScholarPubMed
Ribeiro, P, Gupta, V and El-Sakkary, N (2012) Biogenic amines and the control of neuromuscular signaling in schistosomes. Invertebrate Neuroscience 12, 1328. https://doi.org/10.1007/s10158-012-0132-y.CrossRefGoogle ScholarPubMed
Robb, SM, Ross, E and Sánchez Alvarado, A (2008) SmedGD: The Schmidtea mediterranea genome database. Nucleic Acids Research 36, D599606. https://doi.org/10.1093/nar/gkm684.CrossRefGoogle ScholarPubMed
Romero, AA, Cobb, SA, Collins, JNR, Kliewer, SA, Mangelsdorf, DJ and Collins, JJ III (2021) The Schistosoma mansoni nuclear receptor FTZ-F1 maintains esophageal gland function via transcriptional regulation of meg-8.3. PLOS Pathogens 17(12), e1010140. https://doi.org/10.1371/journal.ppat.1010140.CrossRefGoogle Scholar
Saito, A, Horie, M and Nagase, T (2018) TGF-β signaling in lung health and disease. International Journal of Molecular Science 19, 2460. https://doi.org/10.3390/ijms19082460.CrossRefGoogle ScholarPubMed
Salter, JP, Lim, KC, Hansell, E, Hsieh, I, and McKerrow, JH (2000) Schistosome invasion of human skin and degradation of dermal elastin are mediated by a single serine protease. Journal of Biological Chemistry 275, 3866738673. https://doi.org/10.1074/jbc.M006997200.CrossRefGoogle ScholarPubMed
Samoil, V, Dagenais, M, Ganapathy, V, Aldridge, J, Glebov, A, Jardim, A and Ribeiro, P (2018) Vesicle-based secretion in schistosomes: Analysis of protein and microRNA (miRNA) content of exosome-like vesicles derived from Schistosoma mansoni. Scientific Reports 8, 3286. https://doi.org/10.1038/s41598-018-21587-4.CrossRefGoogle ScholarPubMed
Schote, AB (2007) Nuclear Receptors: Variants and Their Role in Neuro-Endocrine-Immune Regulations. PhD dissertation, University of Trier and the Institute of Immunology, National Laboratory of Health, Luxembourg. Available at https://ubt.opus.hbz-nrw.de/opus45-ubtr/frontdoor/deliver/index/docId/326/file/Thesis_ASF_FINALEDRUCK_HQ.pdf (accessed April 3, 2024).Google Scholar
Severinghaus, AE (1928) Sex studies on Schistosoma japonicum. Quarterly Journal of Microscopical Science 71, 653702. https://doi.org/10.1242/jcs.s2-71.284.653.Google Scholar
Shaker, Y, Samy, N and Ashour, E (2014) Hepatobiliary schistosomiasis. Journal of Clinical and Translational Hepatology 2, 212216. https://doi.org/10.14218/JCTH.2014.00018.Google ScholarPubMed
Shiff, CJ and Dossaji, SF (1991) Ecdysteroids as regulators of host and parasite interactions: A study of interrelationships between Schistosoma mansoni and the host snail, Biomphalaria glabrata. Tropical Medicine and Parasitology 42, 1116.Google ScholarPubMed
Shoemaker, CB, Ramachandran, H, Landa, A, dos Reis, MG and Stein, LD (1992) Alternative splicing of the Schistosoma mansoni gene encoding a homologue of epidermal growth factor receptor. Molecular and Biochemical Parasitology 53, 1732. https://doi.org/10.1016/0166-6851(92)90003-3.CrossRefGoogle ScholarPubMed
Simons, SS Jr, Edwards, DP and Kumar, R (2014) Minireview: Dynamic structures of nuclear hormone receptors: New promises and challenges. Molecular Endocrinology 28, 173182. https://doi.org/10.1210/me.2013-1334.CrossRefGoogle Scholar
Skelly, PJ and Shoemaker, CB (1996) Rapid appearance and asymmetric distribution of glucose transporter SGTP4 at the apical surface of intramammalian-stage Schistosoma mansoni. Proceedings of the National Academy of Sciences of the United States of America 93, 36423646. https://doi.org/10.1073/pnas.93.8.3642.CrossRefGoogle ScholarPubMed
Skelly, PJ, Da’dara, AA, Li, XH, Castro-Borges, W and Wilson, RA (2014) Schistosome feeding and regurgitation. PLOS Pathogens 10, e1004246. https://doi.org/10.1371/journal.ppat.1004246.CrossRefGoogle ScholarPubMed
Stone, WL, Leavitt, L and Varacallo, M (2023) Physiology, growth factor. In StatPearls [Internet]. Treasure Island (FL): StatPearls Publishing. https://www.ncbi.nlm.nih.gov/books/NBK442024/.Google Scholar
Sundaraneedi, MK, Tedla, BA, Eichenberger, RM, Becker, L, Pickering, D, Smout, MJ, Rajan, S, Wangchuk, P, Keene, FR, Loukas, A, Collins, JG and Pearson, MS (2017) Polypyridylruthenium(II) complexes exert antischistosome activity and inhibit parasite acetylcholinesterases. PLOS Neglected Tropical Diseases 11(12), e0006134. https://doi.org/10.1371/journal.pntd.0006134.CrossRefGoogle ScholarPubMed
Taft, AS, Norante, FA and Yoshino, TP (2010) The identification of inhibitors of Schistosoma mansoni miracidial transformation by incorporating a medium-throughput small-molecule screen. Experimental Parasitology 125, 8494. https://doi.org/10.1016/j.exppara.2009.12.021.CrossRefGoogle ScholarPubMed
Taman, A and Ribeiro, P (2009) Investigation of a dopamine receptor in Schistosoma mansoni: functional studies and immunolocalization. Molecular and Biochemical Parasitology 168, 2433. https://doi.org/10.1016/j.molbiopara.2009.06.003.CrossRefGoogle ScholarPubMed
Taman, A and Ribeiro, P (2011a) Glutamate-mediated signaling in Schistosoma mansoni: A novel glutamate receptor is expressed in neurons and the female reproductive tract. Molecular and Biochemical Parasitology 176, 4250. https://doi.org/10.1016/j.molbiopara.2010.12.001.CrossRefGoogle ScholarPubMed
Taman, A and Ribeiro, P (2011b) Characterization of a truncated metabotropic glutamate receptor in a primitive metazoan, the parasitic flatworm Schistosoma mansoni. PLOS ONE 6(11), e27119. https://doi.org/10.1371/journal.pone.0027119.CrossRefGoogle Scholar
The Schistosoma japonicum Genome Sequencing and Functional Analysis Consortium (2009) The Schistosoma japonicum genome reveals features of host-parasite interplay. Nature 460, 345351. https://doi.org/10.1038/nature08140.CrossRefGoogle Scholar
Tobin, AB, Butcher, AJ and Kong, KC (2008) Location, location, location…site-specific GPCR phosphorylation offers a mechanism for cell-type-specific signalling. Trends in Pharmacological Sciences 29, 413420. https://doi.org/10.1016/j.tips.2008.05.006.CrossRefGoogle ScholarPubMed
Tran, EJ, King, MC and Corbett, AH (2014) Macromolecular transport between the nucleus and the cytoplasm: Advances in mechanism and emerging links to disease. Biochimica et Biophysica Acta 1843, 27842795. https://doi.org/10.1016/j.bbamcr.2014.08.003.CrossRefGoogle ScholarPubMed
Vanderstraete, M, Gouignard, N, Ahier, A, Morel, M, Vicogne, J and Dissous, C (2013) The Venus kinase receptor (VKR) family: Structure and evolution. BMC Genomics 14, 361. https://doi.org/10.1186/1471-2164-14-361.CrossRefGoogle ScholarPubMed
Vanderstraete, M, Gouignard, N, Cailliau, K, Morel, M, Hahnel, S, Leutner, S, Beckmann, S, Grevelding, CG and Dissous, C (2014) Venus kinase receptors control reproduction in the platyhelminth parasite Schistosoma mansoni. PLOS Pathogens 10(5), e1004138. https://doi.org/10.1371/journal.ppat.1004138.CrossRefGoogle ScholarPubMed
Verjovski-Almeida, S, DeMarco, R, Martins, EA, Guimarães, PE, Ojopi, EP, Paquola, AC, Piazza, JP, Nishiyama, MY Jr, Kitajima, JP, Adamson, RE, Ashton, PD, Bonaldo, MF, Coulson, PS, Dillon, GP, Farias, LP, Gregorio, SP, Ho, PL, Leite, RA, Malaquias, LC, Marques, RC, Miyasato, PA, Nascimento, AL, Ohlweiler, FP, Reis, EM, Ribeiro, MA, , RG, Stukart, GC, Soares, MB, Gargioni, C, Kawano, T, Rodrigues, V, Madeira, AM, Wilson, RA, Menck, CF, Setubal, JC, Leite, LC and Dias-Neto, E (2003) Transcriptome analysis of the acoelomate human parasite Schistosoma mansoni. Nature Genetics 35, 148157. https://doi.org/10.1038/ng1237.CrossRefGoogle ScholarPubMed
Vicogne, J, Pin, JP, Lardans, V, Capron, M, Noël, C and Dissous, C (2003) An unusual receptor tyrosine kinase of Schistosoma mansoni contains a Venus Flytrap module. Molecular and Biochemical Parasitology 126, 5162. https://doi.org/10.1016/s0166-6851(02)00249-9.CrossRefGoogle ScholarPubMed
Vicogne, J, Cailliau, K, Tulasne, D, Browaeys, E, Yan, YT, Fafeur, V, Vilain, JP, Legrand, D, Trolet, J and Dissous, C (2004) Conservation of epidermal growth factor receptor function in the human parasitic helminth Schistosoma mansoni. Journal of Biological Chemistry 27, 3740737414. https://doi.org/10.1074/jbc.M313738200.CrossRefGoogle Scholar
Vogeler, S, Galloway, TS, Lyons, BP and Bean, TP (2014) The nuclear receptor gene family in the Pacific oyster, Crassostrea gigas, contains a novel subfamily group. BMC Genomics 15, 369. https://doi.org/10.1186/1471-2164-15-369.CrossRefGoogle ScholarPubMed
von Lichtenberg, E (1987) Consequences of infections with schistosomes. In Rollinson, D and Simpson, AJG (eds), The Biology of Schistosomes: From Genes to Latrines. London: Academic Press, 185232.Google Scholar
Walker, AJ (2011) Insights into the functional biology of schistosomes. Parasites and Vectors 4, 203. https://doi.org/10.1186/1756-3305-4-203.CrossRefGoogle ScholarPubMed
Wang, B, Collins, JJ III and Newmark, PA (2013) Functional genomic characterization of neoblast-like stem cells in larval Schistosoma mansoni. Elife 2, e00768. https://doi.org/10.7554/eLife.00768.CrossRefGoogle ScholarPubMed
Wang, S, Luo, X, Zhang, S, Yin, C, Dou, Y and Cai, X (2014) Identification of putative insulin-like peptides and components of insulin signaling pathways in parasitic platyhelminths by the use of genome-wide screening. The FEBS Journal 281, 877893. https://doi.org/10.1111/febs.12655.CrossRefGoogle ScholarPubMed
Wang, J, Chen, R and Collins, JJ III (2019) Systematically improved in vitro culture conditions reveal new insights into the reproductive biology of the human parasite Schistosoma mansoni. PLOS Biology 17(5), e3000254. https://doi.org/10.1371/journal.pbio.3000254.CrossRefGoogle ScholarPubMed
Wang, X, Cheng, S, Chen, X, Zhang, W, Xie, Y, Liu, W, You, Y, Yi, C, Zhu, B, Gu, M, Xu, B, Lu, Y, Wang, J and Hu, W (2022) A metabotropic glutamate receptor affects the growth and development of Schistosoma japonicum. Frontiers in Microbiology 13, 1045490. https://doi.org/10.3389/fmicb.2022.1045490.CrossRefGoogle ScholarPubMed
Weikum, ER, Liu, X and Ortlund, EA (2018) The nuclear receptor superfamily: A structural perspective. Protein Science 27, 18761892. https://doi.org/10.1002/pro.3496.CrossRefGoogle ScholarPubMed
Weis, WI and Kobilka, BK (2018) The molecular basis of G protein-coupled receptor activation. Annual Review of Biochemistry 87, 897919. https://doi.org/10.1146/annurev-biochem-060614-033910.CrossRefGoogle ScholarPubMed
Wendt, GR and Collins, JJ III (2016) Schistosomiasis as a disease of stem cells. Current Opinion in Genetics and Development 40, 95102. https://doi.org/10.1016/j.gde.2016.06.010.Google ScholarPubMed
Wendt, G, Zhao, L, Chen, R, Liu, C, O’Donoghue, AJ, Caffrey, CR, Reese, ML and Collins, JJ (2020) A single-cell RNA-seq atlas of Schistosoma mansoni identifies a key regulator of blood feeding. Science 369, 16441649. https://doi.org/10.1126/science.abb7709.CrossRefGoogle ScholarPubMed
World Health Organization (2024) Schistosomiasis (Bilharzia). Geneva: WHO. Available at https://www.who.int/health-topics/schistosomiasis#tab=tab_1 (accessed April 3, 2024).Google Scholar
WormBase ParaSite (2024) Schistosoma mansoni. Available at https://parasite.wormbase.org/Schistosoma_mansoni_prjea36577/Info/Index/ (accessed December 20, 2024).Google Scholar
Wu, K, Huang, S, Zhao, Y, Umar, A, Chen, H, Yu, Z and Huang, J (2024) Hepatocyte nuclear factor 4 located in different developmental stages in Schistosoma japonicum and involved in important metabolic pathways. Biomedical Journal 13, 100726. https://doi.org/10.1016/j.bj.2024.100726.Google Scholar
Wu, W and LoVerde, PT (2008) Schistosoma mansoni: Identification of SmNR4A, a member of nuclear receptor subfamily 4. Experimental Parasitology 120, 208213. https://doi.org/10.1016/j.exppara.2008.07.005.CrossRefGoogle ScholarPubMed
Wu, W and LoVerde, PT (2011) Nuclear hormone receptors in parasitic helminths. Molecular and Cellular Endocrinology 334, 5666. https://doi.org/10.1016/j.mce.2010.06.011.CrossRefGoogle ScholarPubMed
Wu, W and LoVerde, PT (2019) Nuclear hormone receptors in parasitic Platyhelminths. Molecular and Biochemical Parasitology 233, 111218. https://doi.org/10.1016/j.molbiopara.2019.111218.CrossRefGoogle ScholarPubMed
Wu, W and LoVerde, PT (2021) Identification and evolution of nuclear receptors in Platyhelminths. PLOS One 16(8), e0250750. https://doi.org/10.1371/journal.pone.0250750.CrossRefGoogle ScholarPubMed
Wu, W and LoVerde, PT (2023) Updated knowledge and a proposed nomenclature for nuclear receptors with two DNA binding domains (2DBD-NRs). PLOS One 18(9), e0286107. https://doi.org/10.1371/journal.pone.0286107.CrossRefGoogle Scholar
Wu, W, Niles, EG, El-Sayed, N, Berriman, M and LoVerde, PT (2006) Schistosoma mansoni (Platyhelminthes, Trematoda) nuclear receptors: Sixteen new members and a novel subfamily. Gene 366, 303315. https://doi.org/10.1016/j.gene.2005.09.013.CrossRefGoogle Scholar
Wu, W, Niles, EG, Hirai, H and LoVerde, PT (2007a) Evolution of a novel subfamily of nuclear receptors with members that each contain two DNA binding domains. BMC Evolutionary Biology 7, 27. https://doi.org/10.1186/1471-2148-7-27.CrossRefGoogle ScholarPubMed
Wu, W, Niles, EG and LoVerde, PT (2007b) Thyroid hormone receptor orthologues from invertebrate species with emphasis on Schistosoma mansoni. BMC Evolutionary Biology 7, 150. https://doi.org/10.1186/1471-2148-7-150.CrossRefGoogle ScholarPubMed
Wu, W, Niles, EG, Hirai, H and LoVerde, PT (2007c) Identification and characterization of a nuclear receptor subfamily I member in the Platyhelminth Schistosoma mansoni (SmNR1). The FEBS Journal 274, 390405. https://doi.org/10.1111/j.1742-4658.2006.05587.x.CrossRefGoogle ScholarPubMed
Wu, W, Tak, EY and LoVerde, PT (2008) Schistosoma mansoni: SmE78, a nuclear receptor orthologue of Drosophila ecdysone-induced protein 78. Experimental Parasitology 119, 313318. https://doi.org/10.1016/j.exppara.2008.03.001.CrossRefGoogle ScholarPubMed
Yang, Y, Guo, L, Chen, L, Gong, B, Jia, D and Sun, Q (2023) Nuclear transport proteins: Structure, function, and disease relevance. Signal Transduction and Targeted Therapy 8, 425. https://doi.org/10.1038/s41392-023-01649-4.CrossRefGoogle ScholarPubMed
You, H, Gobert, GN, Jones, MK, Zhang, W and McManus, DP (2011) Signalling pathways and the host-parasite relationship: Putative targets for control interventions against schistosomiasis: Signalling pathways and future anti-schistosome therapies. BioEssays 33, 203214. https://doi.org/10.1002/bies.201000077.CrossRefGoogle ScholarPubMed
You, H, Stephenson, RJ, Gobert, GN and McManus, DP (2014) Revisiting glucose uptake and metabolism in schistosomes: New molecular insights for improved schistosomiasis therapies. Frontiers in Genetics 5, 176. https://doi.org/10.3389/fgene.2014.00176.CrossRefGoogle ScholarPubMed
You, H, Gobert, GN, Cai, P, Mou, R, Nawaratna, S, Fang, G, Villinger, F and McManus, DP (2015) Suppression of the insulin receptors in adult Schistosoma japonicum impacts on parasite growth and development: Further evidence of vaccine potential. PLOS Neglected Tropical Diseases 9, e0003730. https://doi.org/10.1371/journal.pntd.0003730.CrossRefGoogle ScholarPubMed
You, H, Liu, C, Du, X, Nawaratna, S, Rivera, V, Harvie, M, Jones, M and McManus, DP (2018) Suppression of Schistosoma japonicum acetylcholinesterase affects parasite growth and development. International Journal of Molecular Sciences 19, 2426. https://doi.org/10.3390/ijms19082426.CrossRefGoogle ScholarPubMed
Younis, SS, Abou-El-Naga, IF and Radwan, KH (2023) Molluscicidal effect of green synthesized silver nanoparticles using Azadirachta indica on Biomphalaria alexandrina snails and Schistosoma mansoni cercariae. Asian Pacific Journal of Tropical Biomedicine 13, 3544. https://doi.org/10.4103/2221-1691.367688.CrossRefGoogle Scholar
Zamanian, M, Kimber, MJ, McVeigh, P, Carlson, SA, Maule, AG and Day, TA (2011) The repertoire of G protein-coupled receptors in the human parasite Schistosoma mansoni and the model organism Schmidtea mediterranea. BMC Genomics 12, 596. https://doi.org/10.1186/1471-2164-12-596.CrossRefGoogle ScholarPubMed
Zhang, Q, Zhang, S, Chen, J and Xie, Z (2023) The interplay between integrins and immune cells as a regulator in cancer immunology. International Journal of Molecular Sciences 24, 6170. https://doi.org/10.3390/ijms24076170.CrossRefGoogle ScholarPubMed
Zheng, L, Deng, L, Zhong, Y, Wang, Y, Guo, W and Fan, X (2021) Molluscicides against the snail-intermediate host of Schistosoma: A review. Parasitology Research 120, 33553393. https://doi.org/10.1007/s00436-021-07288-4.CrossRefGoogle ScholarPubMed
Figure 0

Table 1. Growth factor receptors

Figure 1

Table 2. Nuclear hormone receptors and nuclear transport receptors

Figure 2

Figure 1. (a) Schematic diagram of a typical nuclear receptor (NR) consisting of A/B, C, D, E, F domains and N and C terminals. The A/B domain includes the activation function 1 (AF-1), C domain is a DNA binding domain (DBD), D domain is a hinge region, E domain contains a ligand binding domain (LBD), and AF-2. (b) Schematic diagram for atypical NRs of S. mansoni containing two DBDs and a single LBD.

Figure 3

Figure 2. Schematic diagram of S. mansoni RNAs transport by XPO5, XPOT, and XPO1. XPO5 (Smp_152800) binds to pre-miRNA, and XPOT (Smp_137650) binds to tRNA directly in a RanGTP dependent manner. Once RNA proteins are exported to the cytoplasm, the RanGTP is converted to GDP resulting in the release of pre-miRNA from the XPO5 and tRNA from the XPOT. Nuclear export of snRNA, rRNA, and mRNA by XPO1 (Smp_124820) requires the cap-binding complex (CBC) protein, a nuclear export signal (NES) containing adaptor protein and RanGTP. The complex transits to the cytoplasm through the nuclear pore complex (NPC) and releases the cargo protein upon the hydrolysis of RanGTP.

Figure 4

Table 3. Neurotransmitter receptors

Figure 5

Figure 3. (a) Schematic diagram of the expected S. mansoni metabotropic glutamate receptor (SmGluR) (Smp_128940) comprising a large N-terminal extracellular ligand binding domain (LBD), a seven-transmembrane (7-TM) anchoring segment, and a C-terminal intracellular domain (ICD) of varying lengths. The LBD is connected to the 7-TM region by a short cysteine-rich domain (CRD). (b)S. mansoni glutamate-binding protein (SmGBP) (Smp_052660) receptor has a conserved ligand binding domain (LBD) but is missing the cysteine-rich domain, the characteristic 7-TM region, and intracellular domain.