Skip to main content

Sex—the most underappreciated variable in research: insights from helminth-infected hosts

Abstract

The sex of a host affects the intensity, prevalence, and severity of helminth infection. In many cases, one sex has been found to be more susceptible than the other, with the prevalence and intensity of helminth infections being generally higher among male than female hosts; however, many exceptions exist. This observed sex bias in parasitism results primarily from ecological, behavioural, and physiological differences between males and females. Complex interactions between these influences modulate the risk of infection. Indeed, an interplay among sex hormones, sex chromosomes, the microbiome and the immune system significantly contributes to the generation of sex bias among helminth-infected hosts. However, sex hormones not only can modulate the course of infection but also can be exploited by the parasites, and helminths appear to have developed molecules and pathways for this purpose. Furthermore, host sex may influence the efficacy of anti-helminth vaccines; however, although little data exist regarding this sex-dependent efficacy, host sex is known to influence the response to vaccines. Despite its importance, host sex is frequently overlooked in parasitological studies. This review focuses on the key contributors to sex bias in the case of helminth infection. The precise nature of the mechanisms/factors determining these sex-specific differences generally remains largely unknown, and this represents an obstacle in the development of control methods. There is an urgent need to identify any protective elements that could be targeted in future therapies to provide optimal disease management with regard to host sex. Hence, more research is needed into the impact of host sex on immunity and protection.

1 Introduction

For decades, animal and human research has demonstrated an overreliance on male subjects and has failed to adequately account for sex differences. As a consequence, our understanding of many diseases and conditions has mostly emerged from studies performed on one sex. However, males and females differ substantially in their susceptibility to viral, bacterial, and parasitic infections, as well as their course of infection, and males and females may respond differently to treatments such as drugs and vaccines [1, 2]. Numerous factors associated with sex affect the response of the immune system to pathogen challenge and the eventual disease outcome. However, the issue of sex disparity has not been widely considered in animal or human research to date (Box 1). As such, relatively little is known about the impact of sex on biology, health and disease.

In addition to the new diseases emerging every few years, long-known diseases, such as helminth infections, remain dangerous and account for many diseases in both humans and animals [3]. Moreover, a high prevalence of helminth infection in farm animals contributes to significant reductions in production levels. Although a variety of helminth species are known to infect a wide range of hosts, they can be categorized into three major groups: nematodes (roundworms), cestodes (tapeworms), and trematodes (flukes); all of these cause infections with significant morbidity and even mortality. Most of the current research regarding parasitic infections has been aimed at developing new control strategies, such as drugs and vaccines. However, the design of new treatments requires a deep understanding of the basic biology of parasites. In vaccine research, an understanding of the immune responses evoked by infection and the identification of protective mechanisms are fundamental prerequisites for further studies; these cannot be investigated properly without the consideration of host sex.

This review examines the influence of host sex in humans and animals during helminth infection. It discusses the influence of sex hormones, sex chromosomes and sex-specific aspects of the microbiome effects on immunity in helminth-exposed hosts. It also considers the impact of sex on treatment efficacy.

2 The impact of host sex on parasitism

Although host sex has already been recognized to have an impact on parasitism, it has not received the attention it deserves. In general, the prevalence and intensity of helminth infections has been found to be higher among male than female hosts, as noted in birds, rodents, ungulates, and humans [4]. This higher resistance among females led to the formulation of the female host supremacy paradigm. However, the model seems to be oversimplified. Male-biased parasitism is not the general rule, as females can be more strongly affected in the case of some infections, or no difference may exist at all between sexes (Table 1). In addition, any present differences may involve differences in the intensity (the number of parasites infecting a host), prevalence (the proportion of infected individuals in population), and severity of infection [5]. Moreover, host sex may influence the course of infection. Indeed, in rats infected with Toxocara canis, different larval infection migratory patterns were observed: a significant increase in larval number was observed in the brain in male rats, while a greater accumulation was noted in the liver in female rats [6].

Table 1 Sex differences in the prevalence and/or intensity of selected helminth infections in respective hosts.

It has been proposed that intrinsic host-related factors predispose one sex to be more susceptible to infection than the other. These host-related factors include physiological influences, such as sex hormones and immunity-related factors, and behavioural influences, which have been associated with differences in susceptibility and exposure, respectively.

Differences in susceptibility are primarily linked to the impact of sex hormones on the immune system. First, while testosterone is crucial in the development of secondary sexual traits in males, it is also a potent immune suppressor. Consequently, the ability to display secondary sexual traits is associated with increased infection risk. Second, under certain conditions, oestrogens can enhance cellular and humoral immune responses in females, thus increasing resistance against infection [7]. These observations are consistent with the immunocompetence handicap model, according to which females will promote their longevity and the survival of their offspring by investing more energy and resources in immune defence than males, while males will invest more into growth and intrasexual competition but will suffer from testosterone-induced immunosuppression [8]. Hence, infection susceptibility is a life-history trait that exists as a trade-off against reproductive effort, which differs between males and females [9].

Differences in exposure may be linked with sex-specific behaviours that result from various sources, including differential habitat use between sexes, higher aggression between males for mating opportunities, the aggregation of one sex and differences in diet among nonhuman animals, with male hosts typically being at a higher risk of infection [10]. As a consequence, males are often more exposed to infective forms of parasites, e.g., the prevalence of Trichuris spp., Varestrongylus spp., and Dictyocaulus spp. is significantly higher in male moose than female moose [11].

Nevertheless, in some circumstances, the behaviour of the male host may favour lower parasitic loads, as seen in Ashworthius sidemi infection in European bison bulls. This can be explained by the fact that European bison males live solitarily or in small groups, while subadults and females with calves tend to aggregate, which increases the likelihood of infection [12]. In addition, sexual dimorphism may also differentially expose the sexes to the parasite. For example, as males are often larger-bodied than females, they may ingest greater amounts of infected prey or may provide a larger area for parasite contact, thus becoming more exposed [10]. Social status also has profound effects on parasite loads in male vertebrates but not in females [13]. Indeed, high-ranking males harbour more parasites than low-ranking males. Increased parasite risk is a cost of high dominance, which is attributed to the priority of access to resources such as food and consequently greater exposure to parasites, as well as the greater mating efforts associated with increased testosterone levels and hence increased susceptibility.

Recently, the resistance/tolerance concept has gained more attention in the context of helminth infection [14]. The concept assumes that hosts can adopt two major lines of defence against infection. The first is to attack parasites directly to reduce worm burdens or eliminate the infection completely (resistance), while the second is to limit the detrimental impact of infection on host health without reducing the parasite load (tolerance). While both strategies aim to maintain host health and improve host fitness, they may have different effects on epidemiology, and their mechanisms may differ. How host tolerance and host resistance affect parasitism remains largely unknown. While the majority of existing research deals with resistance, most studies have overlooked the implications of tolerance, and little research has addressed the impact of host tolerance and resistance against helminth infection in the context of host sex. Research in wild wood mouse populations challenged by multiple helminth species suggests the presence of sex bias in tolerance and resistance: females appear to invest more in immunity but also seem to be more tolerant of parasitic diversity than males [15].

3 The effects of age and gender

There are many confounding factors to consider when analysing sex bias in parasitism, one of which in human studies is gender. While the terms “sex” and “gender” are often treated as synonyms, they cannot be used interchangeably: sex is a biological trait that is determined by specific sex chromosomes (biological construct), whereas gender refers to roles, activities and behaviours that are regulated by cultural and social norms (social construct) [5]. Gender has implications in human studies on the epidemiology of numerous helminth infections, including schistosomiasis. People become infected with Schistosoma spp. through contact with fresh water containing infectious cercariae. In endemic areas, men are known to be more heavily infected with the parasite than women, but they are also at a higher level of exposure through involvement in fishing, which is traditionally a male occupation and carries an increased risk of infection. Such differential exposure between genders may be falsely suggestive of a sex bias and must be carefully considered in epidemiological studies on sex-related differences [16]. Surprisingly, studies on laboratory rodents infected with Schistosoma have demonstrated the opposite trend than in humans, with female mice exhibiting a higher schistosome load than their male counterparts [17]. While the reasons are yet to be elucidated, this difference not only highlights the importance of disparities between laboratory animals and human subjects but also shows that gender may possibly blur physiological host sex effects.

Another confounding factor affecting the likelihood of parasitism is host age. The age of the host should be reported when analysing sex bias when only age-matched groups are compared, as host age may counteract some of the effects of host sex on infection. Unfortunately, data on this topic are scarce. The majority of studies on infection focus either on host sex or host age effects, and rarely both. Age-related differences in infection prevalence may arise from different behaviours and immune statuses associated with age, both of which are affected by the host sex. Typically, helminth infection intensity follows a hump-shaped profile over time, with low parasite load noted at a very young age, reaching maximal values at intermediate age, and then decreasing in older individuals. For example, human schistosomiasis prevalence peaks in school-age children and young adult populations and then gradually declines later in life [18]. In endemic areas, high schistosomiasis prevalence is noted among children due to the large amount of time they spend swimming or bathing in water containing infectious cercariae. Older groups tend to demonstrate less exposure due to age-related changes in behaviour and the development of immunity over time. For similar reasons, children tend to harbour the greatest numbers of intestinal worms when compared with other age groups [19, 20]. However, in some host–parasite systems, the parasites accumulate with time and demonstrate a linear profile of parasite load with age [21,22,23].

Both age and gender may affect disease prevalence, as demonstrated by studies on intestinal parasitic infections in humans in Nepal [24]. At the national level, adults are more likely to be infected than children, and infection rates are higher among girls and young women in rural areas than among their male counterparts. The former observation could be ascribed to the ongoing successful preventive pharmaceutical interventions and educational programmes implemented in schools, and the latter could be due to the fact that school-aged girls in rural areas show low school attendance as a result of gender discrimination and cannot benefit from anti-parasite programmes [24]. Moreover, these girls are enrolled in agricultural work, which increases the risk of infection.

Despite the many subtleties and nuances observed in free-living animals or humans that modify the impact of sex differences on infection, sex bias is also commonly observed under standardized laboratory conditions. This further supports the contention that sex is an important factor affecting health and disease and should be considered a critical variable in infection studies.

The mechanisms underlying sex differences are multifactorial. In addition to environmental and behavioural factors, complex interactions exist between hormonal (different hormone levels), genetic (related to X- and Y-linked genes), and microbiome factors, and these may have a considerable influence on immunity to helminth infection. These sex-specific factors affect the immune response and create sex-dependent differences, and they will be discussed in more detail herein.

4 The effects of sex hormones

Sex hormones include three major groups of steroids, androgens, oestrogens and progestogens, among which testosterone, oestradiol and progesterone are the most important. Beyond reproductive physiology, sex steroids participate in a number of different roles in various nonreproductive tissues, including immune modulation. Most of their effects are mediated via specific receptors that belong to the nuclear receptor superfamily and are hormone-activated transcription factors. These receptors are richly expressed in most cells of the immune system, such as macrophages, dendritic cells, neutrophils, natural killer cells and lymphocytes; upon binding to their cognate hormones, these receptors regulate both the innate and adaptive immune responses. Oestrogens, progestogens and androgens bind specifically to oestrogen receptors (ERs), progesterone receptors (PRs) and androgen receptors (ARs), respectively. Briefly, sex steroids move passively through the membrane of a target cell, interact with their cognate receptor in the cytoplasm, and translocate to the nucleus, where the complex recognizes hormone response elements (HREs) associated with the promoters of target genes to regulate the transcription of genes signalled by the steroid hormone [25, 26]. Moreover, sex steroids can also mediate immediate effects by affecting the regulation of other transcription factors and cytoplasmic signalling events by binding to membrane receptors and influencing the subsequent cross-talk with signalling cascades [27]. In addition, in some circumstances, sex steroid receptors can be activated in the absence of their respective hormones/ligands, thus influencing the immune cell response [28].

As immune cells express ERs, PRs, and ARs, they respond to sex hormones, which can affect various aspects of their function. For example, oestradiol and ER signalling regulate inflammatory pathways in immune cells through putative EREs in the INF-γ promoter, thus evoking Th1-type immunity, while testosterone and AR signalling results in upregulated expression of the tyrosine phosphatase Ptpn1, which decreases Th1 differentiation [29]. A considerable number of genes are regulated by sex hormones, including those associated with immunity [30, 31]; indeed, a genome-wide screening study identified over 70 000 EREs in the human and mouse genomes [32], while the Androgen Responsive Gene Database includes 1785 human genes and 993 mouse genes [33].

Studies of both humans and rodents have shown that sex hormones modulate immune cell functions, which may in turn dictate susceptibility to helminth infection and the course of infection. One study examined the role of sex hormones in the development of Th2 immunity in a sex-biased model of Trichuris muris infection in mice [34]. Enhanced Th2 responses, which are needed for worm expulsion, were mediated by oestradiol, while DHT suppressed Th2 immunity in vitro. This evidence suggests that sex hormones may act as important immunomodulatory factors determining the generation of the Th2 response to intestinal helminth infections. However, host sex and endogenous sex hormone levels are seldom addressed in such studies. It may be speculated that an immunologically permissive environment mediated by testosterone, progesterone or high oestrogen levels may favour helminth infection chronicity, while lower oestrogen levels may promote infection clearance; this will be addressed in more detail below. Clearance may be achieved by shaping the T-cell response towards a Th1, Th2 or T regulatory phenotype. Helminth infection drives the induction of Tregs to induce a state of hyporesponsiveness that favours parasite survival [35]. As Treg function is impacted by testosterone, male hosts may be at greater risk of helminth chronicity. Moreover, the stronger immune responses of females may be associated with more activated innate immune pathways prior to pathogen challenge, as demonstrated by transcriptional studies on the macrophage transcriptome [36]. However, more studies on the impact of host sex on immune cell function in the context of helminth infection are eagerly awaited.

4.1 Impact on innate immunity

Sex steroids modulate innate immune system functions, including the regulation of the inflammatory process and activation of adaptive immunity via antigen presentation. The inflammatory process comprises inflammatory mediator release, phagocytosis, complement system activation and the synthesis of multiple cytokines and chemokines to remove harmful stimuli and start local tissue recovery. All these functions are influenced by sex steroids (Additional file 1).

Oestrogens regulate inflammatory pathways; however, their impact is highly contextual and primarily depends on tissue type, the differential expression of ER subtypes and hormone concentration [37]. Oestradiol typically increases proinflammatory responses at low physiological levels and anti-inflammatory responses at higher concentrations, which are observed during mid- and late pregnancy [38]. Such oestradiol dose dependency has been observed for antigen-presenting cell functions: macrophages/monocytes and dendritic cells (DCs) release prototypic proinflammatory cytokines (e.g., IL-1α, IL-1β) at low oestradiol levels, while proinflammatory IL-1, TNF α and IL-6 secretion is inhibited at higher levels [39], accompanied by a shift towards Th 2 cytokines such as IL-4 and IL-10 [40]. Moreover, oestradiol enhances the expression of TLR-4 on macrophages and promotes the differentiation of inflammatory DCs; this results in the stronger type I INF activity observed in immune cells from females than in immune cells from males [41]. Oestrogen has also been found to play a dual role for NK cells: low hormone levels have stimulatory effects on NK-cell activity, while high levels have suppressive effects [42]. The hormone also influences neutrophil activities, including apoptosis, chemotaxis and NETosis [43, 44]. Again, the effect is dose dependent, as neutrophil apoptosis is more delayed during pregnancy [45].

Androgens exert anti-inflammatory effects on innate immune cells. In contrast to oestradiol, testosterone reduces TLR-4 expression on macrophages, thus directly attenuating proinflammatory responses. Furthermore, testosterone treatment of both macrophages and DCs leads to a reduction in proinflammatory cytokine levels due to the inhibitory effects of AR signalling on transcription factors [41]. Other functions of innate immune cells are also altered: the bactericidal ability of neutrophils has been found to be impaired [46], and monocyte apoptosis has increased [47].

4.2 Impact on adaptive immunity

Sex steroids affect the course of adaptive immunity by modulating not only cell differentiation and number but also the functions of major lymphocyte subsets [48]. The effects of oestrogen are again highly concentration dependent, with low oestradiol levels typically stimulating Th1-type responses and cell-mediated immunity and high concentrations promoting Th2-type responses and humoral immunity [41]. The hormone also enhances the activity of B cells and antibody production [49].

In contrast, androgens generally dampen the adaptive response. Testosterone exerts an inhibitory effect on Th1-cell differentiation, thus contributing to heightened susceptibility to viral infections in males [50]. However, the effect of testosterone on Th2 cell differentiation is not clear [51], as some studies report a promotion of Th2 responses, while others report a suppressive effect or none. Interestingly, androgens have been associated with the suppression of Th2 immunity during helminth infections [34, 52], thus inhibiting the type of response needed to clear the majority of infections. In addition to their impact on T-cell differentiation, androgens also induce regulatory T cells, further dampening immune responses [50].

In conclusion, the literature indicates that oestrogens have both pro- and anti-inflammatory effects on immune compounds, while androgens and progesterone exert suppressive effects [48]. Hence, the different steroid hormone milieu in females and males may yield sex-specific effects.

4.3 Impact of sex hormones during pregnancy

Sex hormones create physiological differences between sexes and are present in different concentrations depending on the sex. Sex steroid levels tend to remain more balanced throughout the lifespan in males, while females experience regular fluctuations throughout the menstrual cycle. These changes in females have a number of advantages from an evolutionary perspective: they contribute to the maintenance of an infection-free environment before ovulation and then create an immunologically permissive environment that favours implantation. Nevertheless, the most drastic changes in hormone levels are observed during gestation. While increased oestrogen and progesterone levels exert immunomodulatory effects that facilitate maternal-foetal tolerance, this is achieved at the expense of impaired immunity. Interestingly, the immune response causes similar biases during pregnancy and most helminth infections, i.e., a shift towards Th2 and regulatory T-cell responses [53]. The impact of high oestrogen levels on immunity has already been mentioned, and progesterone generally suppresses innate immune cell activities. It also inhibits the activation of macrophages and DCs [54] and decreases inflammation by suppressing proinflammatory cytokine production (TNF-α, IFN-γ, and IL-12) and increasing that of anti-inflammatory cytokines such as IL-10 [55]. Furthermore, it suppresses neutrophil, monocyte and NK-cell functions (Additional file 1). Progesterone mediates certain effects on adaptive immunity resulting in a shift from a Th1 to a Th2 response and increases in IL-4, IL-5 and IL-10 production [56]. In addition, high levels of progesterone during pregnancy favour the development of a regulatory T-cell response [57].

With its increased nutritional demands and altered immunity, pregnancy is also associated with an increased risk of acquiring helminth infection, as confirmed by animal and human studies [58, 59]. Helminth infection may be associated with anaemia, preterm birth, impaired foetal growth and pregnancy loss [60]. Again, the outcome is highly contextual; for example, while hookworm infection is associated with a general reduction in female fecundity, fecundity may be increased during Ascaris lumbricoides infection [61].

These contrasting observations may be partially explained by differences in immunity induced by hookworms (mixed Th1/Th2 response) and roundworms (Th2 response). The response evoked during A. lumbricoides infection is favourable for pregnancy, while hookworm infection causes severe iron-deficiency anaemia, which outweighs any effect of immune modulation [61]. However, combined stimulation of Th2 responses by both pregnancy and infection in rats infected with Trichinella spiralis resulted in increased newborn larva (NBL) mortality [62]. It has been reported that progesterone is an inducer of parasiticide activity associated with NBL death [63, 64]. Progesterone was also found to have antiparasiticidal effects against Schistosoma haematobium in female golden hamsters [65]; however, it was found to promote Taenia solium development in vitro [66], and pregnancy increases the prevalence of naturally acquired cysticercosis in rural pigs [67].

The hormonal effects associated with pregnancy are also seen during T. canis and Toxocara cati infection in dogs and cats, respectively. In the case of T. canis, larvae arrested in various tissues of infected bitches become reactivated during gestation, migrating across the placenta and infecting the foetuses [68]. Moreover, T. canis larvae can migrate from mother to neonate via the mammary gland during lactation [69]. While no cases of transplacental larvae transmission have been noted for T. cati, infection can still occur via the lactational route [70]. The reactivation of dormant larvae is mediated by increased progesterone and prolactin levels [68, 69].

4.4 Impact of sex hormones on anti-helminth immunity

When a host encounters a parasite, an interplay begins between host defence mechanisms and parasite survival strategies; this may result in infection clearance or its establishment. This interaction is strongly influenced by host sex hormones. While sex steroids modulate the immune response to infection, they may also directly affect parasite growth, differentiation and reproduction. Nevertheless, the relationship between hormone activity and host susceptibility to helminth infection varies greatly among species and is heavily reliant on the particular parasite-host system.

In the case of nematode infections, males are generally observed to be more susceptible, which is often associated with testosterone levels [71]. Studies on Trichuris muris infection in mice suggest that testosterone demonstrates an inhibitory effect on protective immunity, mainly through a reduction in Th2 cytokine responses [34], and that oestrogens may have protective influences mediated by IL-13 and IL-4 [72, 73]. For Angiostrongylus malaysinensis and Nippostrongylus brasiliensis infections, gonadally intact male rats have higher worm burdens than females or castrated males [74, 75]. Gonadectomy of females does not impact the N. brasiliensis burden [75]. In the case of Strongyloides ratti infection in rats, significantly higher worm burdens are reported in males, while ovariectomy has no effect on parasite load in females. Testosterone treatment increases S. ratti burdens in both males and females [76]. Correspondingly, testosterone levels are positively correlated with worm burden during Haemonchus contortus infection in male lambs [77].

Male mice with higher social rank, and hence increased testosterone levels, are more prone to Heligmosomoides polygyrus infection [78]. However, while male mice are also more susceptible to Brugia malayi challenge than female mice, this is probably more due to the protective effects of the oestrogen-rich environment in the latter rather than the inhibitory effects of testosterone in males [79]. Similarly, testosterone has no significant effect on Trichinella spiralis development, but progesterone and oestradiol treatment inhibits the T. spiralis molting rate in vitro [80].

Sex hormones have also been found to influence the development and survival of cestode helminths; however, in contrast to most nematode infections, increased worm burdens are usually observed among female hosts, such as for Taenia crassiceps infection in gonadally intact mice [81]. Ovariectomy reduces the susceptibility of female mice to parasite challenge, while gonadectomy increases infection intensity among males [82]. In vitro exposure to oestradiol induces cysticerci budding and increases T. crassiceps infective capacity [83], whereas testosterone and dihydrotestosterone reduce parasite survival and impair the excretory system of flame cells, causing parasite intoxication [84, 85]. It has also been suggested that androgens may have a protective role against Taenia solium infection, with in vitro exposure of T. solium cysticerci to testosterone and DHEA inhibiting scolex evagination [83], while progesterone induced the opposite effect [66].

The progression of trematode infection is also affected by sex steroids. Schistosoma mansoni infection is suppressed by elevated testosterone concentrations in male mice [86]. In addition, testosterone appears to have a direct antifecundity influence in adult S. haematobium worms [87].

4.5 Exploitation of sex hormones by helminths

The host endocrine microenvironment can also be exploited by helminths themselves for their own advantage. Such exploitation may include the utilization of receptors, transporters, steroidogenic pathway enzymes and secondary messengers expressed by parasites [88].

4.5.1 Receptors

It has been proposed that parasites have developed molecules analogous to host sex hormone receptors. These bind with the sex steroids of the host, resulting in downstream transcriptional events in the parasite. Indeed, oestrogen receptor-like structures were described in S. mansoni [89], the free-living nematode species Panagrellus redivivus and Caenorhabditis elegans [90], and T. crassiceps [91]. It has been proposed that the interaction between oestrogen receptors and oestrogen-responsive elements leads to the activation of activator protein-1 complex genes, since oestradiol increases the expression of T. crassiceps c-fos and c-jun [83, 92].

Treatment with selective oestrogen receptor modulators has been found to reduce the motility, viability and fertility of adult worms, suggesting that oestrogen receptor-like molecules are present in H. contortus [93]. For S. haematobium, it was demonstrated that testosterone binds with the parasite protein Sh28GST to reduce the fecundity of the parasite [87]. In addition, T. solium expresses a protein related to the progesterone receptor (TsPR), which enables progesterone to have a direct effect on T. solium cysticerci [94]. In teanids, androgens may exert their effects through the nonspecific progesterone receptor membrane component (PGRMC) [95].

Whether these molecules belong to the classic nuclear receptor family remains unclear; although a great number of classic nuclear receptors have been identified in helminths, recent genomic studies suggest this is not the case [96]. It is also possible that sex steroids may passively diffuse through the tegument or may act through membrane nonclassic receptors [95].

4.5.2 Steroidogenic pathway enzymes

There is also evidence that parasites may synthesize steroid hormones from host steroid precursors. In fact, both T. crassiceps and T. solium express steroidogenic enzymes and synthesize steroid hormones [97]. Their cysticerci can transform steroid precursors to androgens. Subsequently, testosterone may be aromatized into oestradiol. Since an oestrogen-rich environment favours teanid growth and development, testosterone production and subsequent transformation into oestradiol may further facilitate the infection progress and may explain the feminization of male mice during chronic infection. Indeed, serum oestrogen concentrations gradually increase following T. crassiceps cysticerci infection in female mice [98], and chronic T. crassiceps infections lead to feminization in males through the overexpression of P-450 [92, 99].

5 The effects of sex chromosomes

Although some of the differences between female and male immunity have been directly attributed to the effects of sex hormones on immune function, sex steroid levels are not sufficient for explaining the disparities observed at different ages (prepubertal, pubertal, post-pubertal), implying that additional mechanisms may be at play. Indeed, many studies indicate that genetic factors also play an important role in sexual dimorphic immunity [100].

In mammals, biological sex is determined by sex chromosomes, with XX denoting females and XY denoting males. The X chromosome carries not only the genes participating in sex determination but also numerous immune-associated genes, including CD40L, CXCR3, FOXP3, TLR7, TLR8, BTK, IRAK-1, and NEMO [101, 102]. While one of the two X chromosomes in females is inactivated by methylation to maintain the same dosage of proteins between the sexes, approximately 15% of X-linked genes escapes the process [103]. As a consequence, some X-encoded immune-associated proteins and factors are overexpressed in females compared to males and contribute to enhanced immune responses in females [2]. Furthermore, the X chromosome inactivation process is random, and hence, females are inherently mosaics composed of cells in which either the maternal or paternal X chromosome is silenced [104]. Such cellular mosaicism is beneficial for females, as it provides them with a greater diversity of responses against the pathogen challenge. Females may also benefit from the maternal transmission of mitochondria, which not only have bioenergetic functions but also are important regulators of immunity [105]. They can regulate the activation, differentiation, survival, and transcription of immune cells [106]. Evolutionary pressures may have forced the selection of mitochondrial alleles that are favourable for females but detrimental for males: the so-called “mother’s curse” [107]. Such a mechanism may negatively influence the disease burden in males by affecting the quality of their immune responses.

Since sex chromosomes affect host immune functions, they may also govern susceptibility to helminth infection and its ultimate outcome. Again, female hosts seem to be better equipped to combat such infection.

6 The effects of the microbiome

In addition to hormonal and genetic factors, both innate and adaptive immunity appear to be influenced by the microbiota inhabiting the body [108]. Moreover, the microbiome composition differs between sexes, as observed in animal and human studies [109, 110]; for example, females have higher levels of Lactobacillaceae, males have higher levels of Ruminococcaceae [111], and females generally have higher microbial diversity and richness than males, which is beneficial for their health [109]. Sex disparities in microbiome composition elicit sex-related immune responses, thus contributing to sex-specific microbiomes [111]. It still remains unknown, however, whether differences in microbiome composition result from different sex steroid levels in males and females or are a cause of the observed sex-specific immunity. This interaction is further complicated by the presence of parasite infection, as it can change the composition of the gut microbiome [112]. For example, Heligmosomoides polygyrus infection increases the abundance of Lactobacillaceae and Enterobacteriaceae in the gut [113, 114]. Moreover, certain Lactobacillus species make the host more susceptible to helminth infection, as demonstrated in studies on Trichuris muris [115], and host sex has been found to alter the response of gut microbiota to cestode infection [116].

7 Efficacy of treatments

Males and females differ not only in their susceptibility to parasitic infections but also in their responsiveness to drugs and vaccines [41]. The sexes are known to react in different ways to pharmacotherapy, with differences in the absorption, metabolism and effectiveness of some medicines being reported [117, 118]. As such, it is highly recommended that sex-specific drug dosing be used to mitigate unnecessary adverse reactions.

Furthermore, although sex influences the course of the immune responses after vaccination, the existence of immunological differences between males and females is rarely considered in vaccine trial design [119]. Sex effects have been reported for many commercially available vaccines [120, 121]. For example, women demonstrate higher humoral responses to measles, hepatitis B, influenza and tetanus vaccines, while men have increased antibody responses to yellow fever, pneumococcal polysaccharide and meningococcal A and C vaccines [122]. Since women are underrepresented in vaccine trials, outcome data are often extrapolated to them from men, thus resulting in inaccurate vaccine dosages [1]. In addition, it has been demonstrated that women vaccinated with a half dose of the influenza vaccine display higher antibody responses than men receiving a full dose [123]. Moreover, due to their higher inflammatory and cellular responses, female recipients tend to experience more adverse effects following vaccination [1].

Helminth infection has a significant influence on the immune response to vaccines and vaccine efficacy [124]. Since most parasites induce systemic immunosuppression in their hosts, protective immune responses to vaccines may be suppressed. The potent regulatory and type 2 immune responses typically elicited by helminth infection may interfere with immunization by vaccines that elicit type 1 immune responses for protection, thus contributing to vaccine failure. Indeed, in endemic areas, helminth-infected children develop poorer influenza-specific responses to vaccines than uninfected groups [125]. Additionally, responsiveness to vaccination against influenza is often suppressed in studies on laboratory rodents, as observed in Litomosoides sigmodontis-infected mice [126].

Helminth-vaccine interactions can result from a qualitative mismatch or agreement in the type of response required to clear the infection and the type of response needed to immunize against the vaccine target. This relationship may be potentially modulated by host sex. As both sexes yield qualitatively and quantitatively different responses, they can demonstrate different responses to vaccines in the context of helminth infection. However, no reliable data exist on the subject, and further studies are needed in this area.

Furthermore, few data exist regarding the sex-dependent efficacy of vaccines against parasitic infections. Few studies deal with malaria [127, 128]. Host sex has also been found to impact the efficacy of vaccines against helminth parasite infections. For instance, host sex influences both vaccine efficacy and immune responses following vaccination and/or infection in laboratory and natural Fasciola hepatica hosts (Table 2).

Table 2 Sex-specific vaccine efficacy.

For example, male rats vaccinated with cDNA encoding F. hepatica phosphoglycerate kinase (cDNA-FhPGK/pCMV) developed marked leucocytosis with higher neutrophil, eosinophil and monocyte responses than females [129]. Additionally, the dynamics of eosinophil and monocyte responses have been found to vary between sexes: increased titres of anti-FhPGK IgG1 and IgG2a correlated with the protective effect of vaccination, but only among female rats [129]. Moreover, during acute and chronic infection, different CD4+ and CD8+ T-cell profiles were noted between males and females in peritoneal fluid and lymph nodes but not in blood [130]. Following cDNA-FhPGK/pCMV vaccination and/or F. hepatica infection, the immune responses of rats were polarized towards Th2/Treg, with lymphocytes isolated from male rats showing higher IL-4 and IL-10 production than females [130]. While lymphocytes isolated from vaccinated and/or infected rats of both sexes had reduced proliferative capacities in response to mitogen (PHA) or vaccine antigen (FhPGK) when compared to those from unvaccinated and uninfected rats, the males demonstrated a considerably greater reduction in proliferative capacity, while the vaccinated females demonstrated greater restored lymphocyte proliferative capacities during chronic fasciolosis [130].

8 Conclusions

The sex of the host affects the fate of helminth infection. Both physiological and behavioural factors play key roles in the differences in susceptibility and exposure reported between sexes. In particular, sex hormones, sex chromosomes and the microbiome have particularly strong influences on the sex bias associated with the immunity of infected hosts. Indeed, the impact of host sex on helminth infection is widespread, and there are multiple examples in which one sex is better protected than the other. However, there is no single overarching mechanism regulating these effects of host sex. In contrast, complex multifaceted interactions exist, and these vary by helminth species and each particular host–parasite system. These complex interactions determine whether the individuals of both sexes are immune, susceptible, or tolerant to helminth infection; further clarification of sex-specific protective factors is needed, particularly the molecular pathways mediating sex-specific differences in infected hosts await identification.

It is likely that inattention to host sex contributes to the lack of success in vaccine development against numerous pathogens, including helminths. Most studies do not provide sex-specific data analysis, which results in an incomplete understanding of immune responses elicited in the two sexes. Research must be undertaken to recognize sex-biasing factors/mechanisms that protect against disease and to support the development of sex-optimized treatments for males and females. If a protective mechanism is identified, it could be augmented or mitigated (as appropriate) to provide optimal disease management. Moreover, even if it seems that infection outcomes are equivalent in males and females, the underlying mechanisms may differ substantially. Excluding one sex may mask discoveries relevant to disease pathogenesis and treatment, while integrating sex into research may increase the likelihood and pace of new discoveries and diminish the risk of extrapolation [131]. It is therefore necessary to intensify and encourage research into the impact of host sex on immunity following helminth infection to provide a better understanding of how the immune system functions.

Availability of data and materials

The data supporting the conclusions of this article are included within the article.

References

  1. Ruggierii A, Anticoli S, D’ambrosio A, Giordani L, Mora M (2016) The influence of sex and gender on immunity, infection and vaccination. Ann Ist Super Sanita 52:198–204. https://doi.org/10.4415/ANN_16_02_11

    Article  Google Scholar 

  2. Schurz H, Salie M, Tromp G, Hoal EG, Kinnear CJ, Möller M (2019) The X chromosome and sex-specific effects in infectious disease susceptibility. Hum Genom 13:2. https://doi.org/10.1186/S40246-018-0185-Z

    Article  Google Scholar 

  3. Hotez PJ, Brindley PJ, Bethony JM, King CH, Pearce EJ, Jacobson J (2008) Helminth infections: the great neglected tropical diseases. J Clin Invest 118:1311–1321. https://doi.org/10.1172/JCI34261

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Oliver-Guimerá A, Martínez-Carrasco C, Tvarijonaviciute A, Ruiz de Ybáñez MR, Martínez-Guijosa J, López-Olvera JR, Fernandez-Aguilar X, Colom-Cadena A, Mentaberre G, Velarde R, Gassó D, Garel M, Rossi L, Lavin S, Serrano E (2017) The physiological cost of male-biased parasitism in a nearly monomorphic mammal. Parasites Vectors 10:200. https://doi.org/10.1186/S13071-017-2060-5

    Article  PubMed  PubMed Central  Google Scholar 

  5. vom Steeg LG, Klein SL (2016) SeXX matters in infectious disease pathogenesis. PLoS Pathog 12:e1005374. https://doi.org/10.1371/JOURNAL.PPAT.1005374

    Article  Google Scholar 

  6. Yazawa F, Chieffi P, Lescano S, Fonseca G, dos Santos SV, (2018) Effect of sex on Toxocara larval migration to the cerebellum during experimental infection of Rattus norvegicus. J Trop Pathol 47:111–115

    Google Scholar 

  7. Klein SL (2004) Hormonal and immunological mechanisms mediating sex differences in parasite infection. Parasite Immunol 26:247–264. https://doi.org/10.1111/J.0141-9838.2004.00710.X

    Article  CAS  PubMed  Google Scholar 

  8. Córdoba-Aguilar A, Munguía-Steyer R (2013) The sicker sex: understanding male biases in parasitic infection, resource allocation and fitness. PLoS One 8:e76246. https://doi.org/10.1371/JOURNAL.PONE.0076246

    Article  PubMed  PubMed Central  Google Scholar 

  9. Stoehr AM, Kokko H (2006) Sexual dimorphism in immunocompetence: what does life-history theory predict? Behav Ecol 17:751–756. https://doi.org/10.1093/BEHECO/ARK018

    Article  Google Scholar 

  10. Brown DS, Symondson WOC (2014) Sex and age-biased nematode prevalence in reptiles. Mol Ecol 23:3890–3899. https://doi.org/10.1111/MEC.12688

    Article  PubMed  Google Scholar 

  11. Filip-Hutsch K, Czopowicz M, Świsłocka M, Ratkiewicz M, Borkowska A, Kowalczyk R, Demiaszkiewicz AW (2020) Patterns of parasite eggs, oocysts and larvae shedding by moose in the Biebrza marshland (NE Poland). Int J Parasitol Parasites Wildl 11:191–197. https://doi.org/10.1016/J.IJPPAW.2020.02.007

    Article  PubMed  PubMed Central  Google Scholar 

  12. Krasińska M, Krasiński ZABA (2000) Factors affecting the variability in home range size and distribution in European bison in the Polish and Belarussian parts of the Białowieża Forest. Acta Theriol 45:321–334

    Article  Google Scholar 

  13. Habig B, Doellman MM, Woods K, Olansen J, Archie EA (2018) Social status and parasitism in male and female vertebrates: a meta-analysis. Sci Rep 8:3629. https://doi.org/10.1038/S41598-018-21994-7

    Article  PubMed  PubMed Central  Google Scholar 

  14. Råberg L, Graham AL, Read AF (2009) Decomposing health: tolerance and resistance to parasites in animals. Philos Trans R Soc Lond B Biol Sci 364:37–49. https://doi.org/10.1098/RSTB.2008.0184

    Article  PubMed  Google Scholar 

  15. Bordes F, Ponlet N, Goüy De Bellocq J, Ribas A, Krasnov BR, Morand S (2012) Is there sex-biased resistance and tolerance in Mediterranean wood mouse (Apodemus sylvaticus) populations facing multiple helminth infections? Oecologia 170:123–135. https://doi.org/10.1007/s00442-012-2300-5

    Article  PubMed  Google Scholar 

  16. Bernin H, Lotter H (2014) Sex bias in the outcome of human tropical infectious diseases: influence of steroid hormones. J Infect Dis 209:S107–S113. https://doi.org/10.1093/INFDIS/JIT610

    Article  CAS  PubMed  Google Scholar 

  17. Boissier J, Chlichlia K, Digon Y, Ruppel A, Moné H (2003) Preliminary study on sex-related inflammatory reactions in mice infected with Schistosoma mansoni. Parasitol Res 91:144–150. https://doi.org/10.1007/S00436-003-0943-1

    Article  CAS  PubMed  Google Scholar 

  18. LoVerde PT (2019) Schistosomiasis. Adv Exp Med Biol 1154:45–70. https://doi.org/10.1007/978-3-030-18616-6_3

    Article  PubMed  Google Scholar 

  19. Ngonjo T, Okoyo C, Andove J, Simiyu E, Lelo AE, Kabiru E, Kihara J, Mwandawiro C (2016) Current status of soil-transmitted helminths among school children in Kakamega County, Western Kenya. J Parasitol Res 2016:7680124. https://doi.org/10.1155/2016/7680124

    Article  PubMed  PubMed Central  Google Scholar 

  20. Blackwell AD, Gurven MD, Sugiyama LS, Madimenos FC, Liebert MA, Martin MA, Kaplan HS, Snodgrass JJ (2011) Evidence for a peak shift in a humoral response to helminths: age profiles of IgE in the Shuar of Ecuador, the Tsimane of Bolivia, and the US NHANES. PLoS Negl Trop Dis 5:e1218. https://doi.org/10.1371/JOURNAL.PNTD.0001218

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Burlet P, Deplazes P, Hegglin D (2011) Age, season and spatio-temporal factors affecting the prevalence of Echinococcus multilocularis and Taenia taeniaeformis in Arvicola terrestris. Parasites Vectors 4:6. https://doi.org/10.1186/1756-3305-4-6

    Article  PubMed  PubMed Central  Google Scholar 

  22. Theis JH, Schwab RG (1992) Seasonal prevalence of Taenia taeniaeformis: relationship to age, sex, reproduction and abundance of an intermediate host (Peromyscus maniculatus). J Wildl Dis 28:42–50. https://doi.org/10.7589/0090-3558-28.1.42

    Article  CAS  PubMed  Google Scholar 

  23. Bellay S, Oda FH, Almeida-Neto M, de Oliveira EF, Takemoto RM, Balbuena JA (2020) Host age predicts parasite occurrence, richness, and nested infracommunities in a pilot whale-helminth network. Parasitol Res 119:2237–2244. https://doi.org/10.1007/S00436-020-06716-1

    Article  PubMed  Google Scholar 

  24. Bertoncello C, Amoruso I, Moscardino U, Fonzo M, Maharjan M, Buja A, Baldo V, Cocchio S, Baldovin T (2021) Sex-biased prevalence of intestinal parasitic infections and gender inequality in rural Nepal. Int J Infect Dis 109:148–154. https://doi.org/10.1016/J.IJID.2021.06.041

    Article  PubMed  Google Scholar 

  25. Moulton VR (2018) Sex hormones in acquired immunity and autoimmune disease. Front Immunol 9:2279. https://doi.org/10.3389/FIMMU.2018.02279

    Article  PubMed  PubMed Central  Google Scholar 

  26. Ben-Batalla I, Vargas-Delgado ME, von Amsberg G, Janning M, Loges S (2020) Influence of androgens on immunity to self and foreign: effects on immunity and cancer. Front Immunol 11:1184. https://doi.org/10.3389/FIMMU.2020.01184

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Fuentes N, Silveyra P (2019) Estrogen receptor signaling mechanisms. Adv Protein Chem Struct Biol 116:135–170. https://doi.org/10.1016/BS.APCSB.2019.01.001

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Bennesch MA, Picard D (2015) Minireview: tipping the balance: ligand-independent activation of steroid receptors. Mol Endocrinol 29:349–363. https://doi.org/10.1210/ME.2014-1315

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Kissick HT, Sanda MG, Dunn LK, Pellegrini KL, On ST, Noel JK, Arredouani MS (2014) Androgens alter T-cell immunity by inhibiting T-helper 1 differentiation. Proc Natl Acad Sci USA 111:9887–9892. https://doi.org/10.1073/PNAS.1402468111

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Snyder EM, Small CL, Li Y, Griswold MD (2009) Regulation of gene expression by estrogen and testosterone in the proximal mouse reproductive tract. Biol Reprod 81:707–716. https://doi.org/10.1095/BIOLREPROD.109.079053

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Tang S, Han H, Bajic VB (2004) ERGDB: estrogen responsive genes database. Nucleic Acids Res 32:D533-536. https://doi.org/10.1093/NAR/GKH083

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Bourdeau V, Deschênes J, Métivier R, Nagai Y, Nguyen D, Bretschneider N, Gannon F, White JH, Mader S (2004) Genome-wide identification of high-affinity estrogen response elements in human and mouse. Mol Endocrinol 18:1411–1427. https://doi.org/10.1210/ME.2003-0441

    Article  CAS  PubMed  Google Scholar 

  33. Jiang M, Ma Y, Chen C, Fu X, Yang S, Li X, Yu G, Mao Y, Xie Y, Li Y (2009) Androgen-responsive gene database: integrated knowledge on androgen-responsive genes. Mol Endocrinol 23:1927–1933. https://doi.org/10.1210/ME.2009-0103

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Hepworth MR, Hardman MJ, Grencis RK (2010) The role of sex hormones in the development of Th2 immunity in a gender-biased model of Trichuris muris infection. Eur J Immunol 40:406–416. https://doi.org/10.1002/EJI.200939589

    Article  CAS  PubMed  Google Scholar 

  35. White MPJ, McManus CM, Maizels RM (2020) Regulatory T-cells in helminth infection: induction, function and therapeutic potential. Immunology 160:248–260. https://doi.org/10.1111/imm.13190

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Gal-Oz ST, Maier B, Yoshida H, Seddu K, Elbaz N, Czysz C, Zuk O, Stranger BE, Ner-Gaon H, Shay T (2019) ImmGen report: sexual dimorphism in the immune system transcriptome. Nat Commun 10:4295. https://doi.org/10.1038/s41467-019-12348-6

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Kovats S (2015) Estrogen receptors regulate innate immune cells and signaling pathways. Cell Immunol 294:63–69. https://doi.org/10.1016/J.CELLIMM.2015.01.018

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Gabriel G, Arck PC (2014) Sex, immunity and influenza. J Infect Dis 209:S93–S99. https://doi.org/10.1093/infdis/jiu020

    Article  PubMed  Google Scholar 

  39. Zhang X, Wang L, Zhang H, Guo D, Qiao Z, Qiao J (2001) Estrogen inhibits lipopolysaccharide-induced tumor necrosis factor-alpha release from murine macrophages. Methods Find Exp Clin Pharmacol 23:169–173. https://doi.org/10.1358/MF.2001.23.4.634640

    Article  CAS  PubMed  Google Scholar 

  40. Liu HY, Buenafe AC, Matejuk A, Ito A, Zamora A, Dwyer J, Vandenbark AA, Offner H (2002) Estrogen inhibition of EAE involves effects on dendritic cell function. J Neurosci Res 70:238–248. https://doi.org/10.1002/JNR.10409

    Article  CAS  PubMed  Google Scholar 

  41. Klein SL, Flanagan KL (2016) Sex differences in immune responses. Nat Rev Immunol 16:626–638. https://doi.org/10.1038/NRI.2016.90

    Article  CAS  PubMed  Google Scholar 

  42. Straub RH (2007) The complex role of estrogens in inflammation. Endocr Rev 28:521–574. https://doi.org/10.1210/ER.2007-0001

    Article  CAS  PubMed  Google Scholar 

  43. Sowemimo OA (2009) The prevalence and intensity of gastrointestinal parasites of dogs in Ile-Ife, Nigeria. J Helminthol 83:27–31. https://doi.org/10.1017/S0022149X08067229

    Article  PubMed  Google Scholar 

  44. Shepherd R, Cheung AS, Pang K, Saffery R, Novakovic B (2021) Sexual dimorphism in innate immunity: the role of sex hormones and epigenetics. Front Immunol 11:3559. https://doi.org/10.3389/FIMMU.2020.604000/BIBTEX

    Article  Google Scholar 

  45. Molloy EJ, O’Neill AJ, Grantham JJ, Sheridan-Pereira M, Fitzpatrick JM, Webb DW, Watson RW (2003) Sex-specific alterations in neutrophil apoptosis: the role of estradiol and progesterone. Blood 102:2653–2659. https://doi.org/10.1182/BLOOD-2003-02-0649

    Article  CAS  PubMed  Google Scholar 

  46. Scalerandi MV, Peinetti N, Leimgruber C, Rubio MMC, Nicola JP, Menezes GB, Maldonado CA, Quintar A (2018) Inefficient N2-like neutrophils are promoted by androgens during infection. Front Immunol 9:1980. https://doi.org/10.3389/FIMMU.2018.01980

    Article  PubMed  PubMed Central  Google Scholar 

  47. Cutolo M, Capellino S, Montagna P, Ghiorzo P, Sulli A, Villaggio B (2005) Sex hormone modulation of cell growth and apoptosis of the human monocytic/macrophage cell line. Arthritis Res Ther 7:R1124–R1132. https://doi.org/10.1186/AR1791

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Bereshchenko O, Bruscoli S, Riccardi C (2018) Glucocorticoids, sex hormones, and immunity. Front Immunol 9:1332. https://doi.org/10.3389/FIMMU.2018.01332

    Article  PubMed  PubMed Central  Google Scholar 

  49. Bernardi AI, Andersson A, Grahnemo L, Nurkkala-Karlsson M, Ohlsson C, Carlsten H, Islander U (2014) Effects of lasofoxifene and bazedoxifene on B cell development and function. Immun Inflamm Dis 2:214–225. https://doi.org/10.1002/IID3.37

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Trigunaite A, Dimo J, Jørgensen TN (2015) Suppressive effects of androgens on the immune system. Cell Immunol 294:87–94. https://doi.org/10.1016/J.CELLIMM.2015.02.004

    Article  CAS  PubMed  Google Scholar 

  51. Henze L, Schwinge D, Schramm C (2020) The effects of androgens on T cells: clues to female predominance in autoimmune liver diseases? Front Immunol 11:1567. https://doi.org/10.3389/FIMMU.2020.01567/BIBTEX

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Pearce EJ, Kane CM, Sun J, Taylor JJ, McKee AS, Cervi L (2004) Th2 response polarization during infection with the helminth parasite Schistosoma mansoni. Immunol Rev 201:117–126. https://doi.org/10.1111/J.0105-2896.2004.00187.X

    Article  CAS  PubMed  Google Scholar 

  53. Persson G, Ekmann JR, Hviid TVF (2019) Reflections upon immunological mechanisms involved in fertility, pregnancy and parasite infections. J Reprod Immunol 136:102610. https://doi.org/10.1016/J.JRI.2019.08.001

    Article  CAS  PubMed  Google Scholar 

  54. Jones LA, Kreem S, Shweash M, Paul A, Alexander J, Roberts CW (2010) Differential modulation of TLR3- and TLR4-mediated dendritic cell maturation and function by progesterone. J Immunol 185:4525–4534. https://doi.org/10.4049/JIMMUNOL.0901155

    Article  CAS  PubMed  Google Scholar 

  55. Hall OJ, Klein SL (2017) Progesterone-based compounds affect immune responses and susceptibility to infections at diverse mucosal sites. Mucosal Immunol 10:1097–1107. https://doi.org/10.1038/MI.2017.35

    Article  CAS  PubMed  Google Scholar 

  56. Miyaura H, Iwata M (2002) Direct and indirect inhibition of Th1 development by progesterone and glucocorticoids. J Immunol 168:1087–1094. https://doi.org/10.4049/JIMMUNOL.168.3.1087

    Article  CAS  PubMed  Google Scholar 

  57. Mao G, Wang J, Kang Y, Tai P, Wen J, Zou Q, Li G, Ouyang H, Xia G, Wang B (2010) Progesterone increases systemic and local uterine proportions of CD4+CD25+ Treg cells during midterm pregnancy in mice. Endocrinology 151:5477–5488. https://doi.org/10.1210/EN.2010-0426

    Article  CAS  PubMed  Google Scholar 

  58. Pelletier F, Ann Page K, Ostiguy T, Festa-Bianchet Pelletier M (2005) Fecal counts of lungworm larvae and reproductive effort in bighorn sheep, Ovis canadensis. Oikos 110:473–480

    Article  Google Scholar 

  59. González-Fernández D, Koski KG, Sinisterra OT, Del Carmen PE, Murillo E, Scott ME (2015) Interactions among urogenital, intestinal, skin, and oral infections in pregnant and lactating Panamanian Ngäbe women: a neglected public health challenge. Am J Trop Med Hyg 92:1100–1110. https://doi.org/10.4269/AJTMH.14-0547

    Article  PubMed  PubMed Central  Google Scholar 

  60. Blackwell AD (2016) Helminth infection during pregnancy: insights from evolutionary ecology. Int J Womens Health 8:651. https://doi.org/10.2147/IJWH.S103529

    Article  PubMed  PubMed Central  Google Scholar 

  61. Blackwell AD, Tamayo MA, Beheim B, Trumble BC, Stieglitz J, Hooper PL, Martin M, Kaplan H, Gurven M (2015) Helminth infection, fecundity, and age of first pregnancy in women. Science 350:970–972. https://doi.org/10.1126/science.aac7902

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Nuñez G, Gentile T, Calcagno M, Venturiello S (2002) Increased parasiticide activity against Trichinella spiralis newborn larvae during pregnancy. Parasitol Res 88:661–667. https://doi.org/10.1007/S00436-002-0599-2

    Article  PubMed  Google Scholar 

  63. Nuñez GG, Gentile T, Costantino SN, Sarchi MI, Venturiello SM (2005) In vitro and in vivo effects of progesterone on Trichinella spiralis newborn larvae. Parasitology 131:255–259. https://doi.org/10.1017/S0031182005007468

    Article  CAS  PubMed  Google Scholar 

  64. Hlaka L, Chitanga S, Masola B, Mukaratirwa S (2017) Host pregnancy influences the establishment of Trichinella zimbabwensis in Balb C mice. J Parasit Dis 41:799–804. https://doi.org/10.1007/S12639-017-0891-9

    Article  PubMed  PubMed Central  Google Scholar 

  65. Soliman MFM, Ibrahim MM (2005) Antischistosomal action of atorvastatin alone and concurrently with medroxyprogesterone acetate on Schistosoma haematobium harboured in hamster: surface ultrastructure and parasitological study. Acta Trop 93:1–9. https://doi.org/10.1016/J.ACTATROPICA.2004.08.006

    Article  CAS  PubMed  Google Scholar 

  66. Escobedo G, Camacho-Arroyo I, Hernández-Hernández OT, Ostoa-Saloma P, García-Varela M, Morales-Montor J (2010) Progesterone induces scolex evagination of the human parasite Taenia solium: evolutionary implications to the host–parasite relationship. J Biomed Biotechnol 2010:591079. https://doi.org/10.1155/2010/591079

    Article  CAS  PubMed  Google Scholar 

  67. Morales J, Velasco T, Tovar V, Fragoso G, Fleury A, Beltrán C, Villalobos N, Aluja A, Rodarte LF, Sciutto E, Larralde C (2002) Castration and pregnancy of rural pigs significantly increase the prevalence of naturally acquired Taenia solium cysticercosis. Vet Parasitol 108:41–48. https://doi.org/10.1016/S0304-4017(02)00168-1

    Article  CAS  PubMed  Google Scholar 

  68. Chávez-Güitrón LE, Morales-Montor J, Nava-Castro KE, Ramírez-Álvarez H, Moreno-Mendoza NA, Prado-Ochoa MG, Muñoz-Guzmán MA, Alba-Hurtado F (2019) Progesterone in vitro increases growth, motility and progesterone receptor expression in third stage larvae of Toxocara canis. Exp Parasitol 198:1–6. https://doi.org/10.1016/J.EXPPARA.2019.01.001

    Article  PubMed  Google Scholar 

  69. Jin Z, Akao N, Ohta N (2008) Prolactin evokes lactational transmission of larvae in mice infected with Toxocara canis. Parasitol Int 57:495–498. https://doi.org/10.1016/J.PARINT.2008.06.006

    Article  CAS  PubMed  Google Scholar 

  70. Maciag L, Morgan ER, Holland C (2022) Toxocara: time to let cati ‘out of the bag.’ Trends Parasitol 38:280–289. https://doi.org/10.1016/J.PT.2021.12.006

    Article  PubMed  Google Scholar 

  71. Roberts CW, Horsnell WGC, Roberts CW, Horsnell WGC (2015) Effects of sex and maternal immunity on protozoan and helminth infections. In: Klein S, Roberts C (eds) Sex and gender differences in infection and treatments for infectious diseases. Springer, Cham, pp 361–388. https://doi.org/10.1007/978-3-319-16438-0_13

    Chapter  Google Scholar 

  72. Bancroft AJ, McKenzie ANJ, Grencis RK (1998) A critical role for IL-13 in resistance to intestinal nematode infection. J Immunol 160:3453–3461

    CAS  PubMed  Google Scholar 

  73. Bancroft AJ, Artis D, Donaldson DD, Sypek JP, Grencis RK (2000) Gastrointestinal nematode expulsion in IL-4 knockout mice is IL-13 dependent. Eur J Immunol 30:2083–2091. https://doi.org/10.1002/1521-4141

    Article  CAS  PubMed  Google Scholar 

  74. Kamis AB, Ahmad RA, Badrul-Munir MZ (1992) Worm burden and leukocyte response in Angiostrongylus malaysiensis-infected rats: the influence of testosterone. Parasitol Res 78:388–391. https://doi.org/10.1007/BF00931693

    Article  CAS  PubMed  Google Scholar 

  75. Tiuria R, Horii Y, Tateyama S, Tsuchiya K, Nawa Y (1994) The Indian soft-furred rat, Millardia meltada, a new host for Nippostrongylus brasiliensis, showing androgen-dependent sex difference in intestinal mucosal defence. Int J Parasitol 24:1055–1057. https://doi.org/10.1016/0020-7519(94)90170-8

    Article  CAS  PubMed  Google Scholar 

  76. Kiyota M, Korenaga M, Nawa Y, Kotani M (1984) Effect of androgen on the expression of the sex difference in susceptibility to infection with Strongyloides ratti in C57BL/6 mice. Aust J Exp Biol Med Sci 62:607–618. https://doi.org/10.1038/ICB.1984.58

    Article  CAS  PubMed  Google Scholar 

  77. Gauly M, Schackert M, Hoffmann B, Erhardt G (2006) Influence of sex on the resistance of sheep lambs to an experimental Haemonchus contortus infection. Dtsch Tierarztl Wochenschr 113:178–181

    CAS  PubMed  Google Scholar 

  78. Barnard CJ, Behnke JM, Gage AR, Brown H, Smithurst PR (1998) The role of parasite-induced immunodepression, rank and social environment in the modulation of behaviour and hormone concentration in male laboratory mice (Mus musculus). Proc Biol Sci 265:693–701. https://doi.org/10.1098/RSPB.1998.0349

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Rajan TV, Nelson FK, Shultz LD, Shultz KL, Beamer WG, Yates J, Greiner DL (1994) Influence of gonadal steroids on susceptibility to Brugia malayi in scid mice. Acta Trop 56:307–314. https://doi.org/10.1016/0001-706X(94)90102-3

    Article  CAS  PubMed  Google Scholar 

  80. Hernández-Bello R, Ramirez-Nieto R, Muñiz-Hernández S, Nava-Castro K, Pavón L, Sánchez-Acosta AG, Morales-Montor J (2011) Sex steroids effects on the molting process of the helminth human parasite Trichinella spiralis. J Biomed Biotechnol 2011:625380. https://doi.org/10.1155/2011/625380

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Fragoso G, Meneses G, Sciutto E, Fleury A, Larralde C (2008) Preferential growth of Taenia crassiceps cysticerci in female mice holds across several laboratory mice strains and parasite lines. J Parasitol 94:551–553. https://doi.org/10.1645/GE-1287.1

    Article  CAS  PubMed  Google Scholar 

  82. Huerta L, Terrazas L, Sciutto E, Larralde C (1992) Immunological mediation of gonadal effects on experimental murine cysticercosis caused by Taenia crassiceps metacestodes. J Parasitol 78:471–476

    Article  CAS  PubMed  Google Scholar 

  83. Escobedo G, Larralde C, Chavarria A, Cerbón MA, Morales-Montor J (2004) Molecular mechanisms involved in the differential effects of sex steroids on the reproduction and infectivity of Taenia crassiceps. J Parasitol 90:1235–1244. https://doi.org/10.1645/GE-297R

    Article  CAS  PubMed  Google Scholar 

  84. Ambrosio JR, Ostoa-Saloma P, Palacios-Arreola MI, Ruíz-Rosado A, Sánchez-Orellana PL, Reynoso-Ducoing O, Nava-Castro KE, Martínez-Velázquez N, Escobedo G, Ibarra-Coronado EG, Valverde-Islas L, Morales-Montor J (2014) Oestradiol and progesterone differentially alter cytoskeletal protein expression and flame cell morphology in Taenia crassiceps. Int J Parasitol 44:687–696. https://doi.org/10.1016/J.IJPARA.2014.04.004

    Article  CAS  PubMed  Google Scholar 

  85. Ambrosio JR, Valverde-Islas L, Nava-Castro KE, Palacios-Arreola MI, Ostoa-Saloma P, Reynoso-Ducoing O, Escobedo G, Ruiz-Rosado A, Domingez-Ramirez L, Morales-Montor J (2015) Androgens exert a cysticidal effect upon Taenia crassiceps by disrupting flame cell morphology and function. PLoS One 10:e0127928. https://doi.org/10.1371/JOURNAL.PONE.0127928

    Article  PubMed  PubMed Central  Google Scholar 

  86. Nakazawa M, Fantappie MR, Freeman GL Jr, Eloi-Santos S, Olsen NJ, Kovacs WJ, Secor WE, Colley DG (1997) Schistosoma mansoni: susceptibility differences between male and female mice can be mediated by testosterone during early infection. Exp Parasitol 85:233–240. https://doi.org/10.1006/EXPR.1997.4148

    Article  CAS  PubMed  Google Scholar 

  87. Remoué F, Mani JC, Pugnière M, Schacht AM, Capron A, Riveau G (2002) Functional specific binding of testosterone to Schistosoma haematobium 28-kilodalton glutathione S-transferase. Infect Immun 70:601–605. https://doi.org/10.1128/IAI.70.2.601-605.2002

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Escobedo G, Roberts CW, Carrero JC, Morales-Montor J (2005) Parasite regulation by host hormones: an old mechanism of host exploitation? Trends Parasitol 21:588–593. https://doi.org/10.1016/J.PT.2005.09.013

    Article  CAS  PubMed  Google Scholar 

  89. Barrabes A, Goma-Mouanda J, Reynouard F, Combescot C (1986) 17 beta-estradiol receptors in Schistosoma mansoni. Contribution to the explanation of the protective power of this hormone in Schistosoma mansoni bilharziasis in the mouse. Preliminary study. Ann Parasitol Hum Comp 61:637–641. https://doi.org/10.1051/PARASITE/1986616637

    Article  CAS  PubMed  Google Scholar 

  90. Hood TE, Calabrese EJ, Zuckerman BM (2000) Detection of an estrogen receptor in two nematode species and inhibition of binding and development by environmental chemicals. Ecotoxicol Environ Saf 47:74–81. https://doi.org/10.1006/EESA.2000.1917

    Article  CAS  PubMed  Google Scholar 

  91. Ibarra-Coronado EG, Escobedo G, Nava-Castro K, Jesús Ramses CR, Hernández-Bello R, García-Varela M, Ambrosio JR, Reynoso-Ducoing O, Fonseca-Liñán R, Ortega-Pierres G, Pavón L, Hernández ME, Morales-Montor J (2011) A helminth cestode parasite express an estrogen-binding protein resembling a classic nuclear estrogen receptor. Steroids 76:1149–1159. https://doi.org/10.1016/J.STEROIDS.2011.05.003

    Article  CAS  PubMed  Google Scholar 

  92. Morales-Montor J, Escobedo G, Rodriguez-Dorantes M, Téllez-Ascencio N, Cerbón MA, Larralde C (2004) Differential expression of AP-1 transcription factor genes c-fos and c-jun in the helminth parasites Taenia crassiceps and Taenia solium. Parasitology 129:233–243. https://doi.org/10.1017/S0031182004005529

    Article  CAS  PubMed  Google Scholar 

  93. Preston S, Luo J, Zhang Y, Jabbar A, Crawford S, Baell J, Hofmann A, Hu M, Gasser RB (2016) Selenophene and thiophene-core estrogen receptor ligands that inhibit motility and development of parasitic stages of Haemonchus contortus. Parasites Vectors 9:346. https://doi.org/10.1186/S13071-016-1612-4

    Article  PubMed  PubMed Central  Google Scholar 

  94. Aguilar-Díaz H, Nava-Castro KE, Escobedo G, Domínguez-Ramírez L, García-Varela M, Del Río-Araiza VH, Palacios-Arreola MI, Morales-Montor J (2018) A novel progesterone receptor membrane component (PGRMC) in the human and swine parasite Taenia solium: implications to the host–parasite relationship. Parasites Vectors 11:161. https://doi.org/10.1186/S13071-018-2703-1

    Article  PubMed  PubMed Central  Google Scholar 

  95. Ambrosio JR, Palacios-Arreola MI, Ríos-Valencia DG, Reynoso-Ducoing O, Nava-Castro KE, Ostoa-Saloma P, Morales-Montor J (2019) Proteomic profile associated with cell death induced by androgens in Taenia crassiceps cysticerci: proposed interactome. J Helminthol 93:539–547. https://doi.org/10.1017/S0022149X18000706

    Article  CAS  PubMed  Google Scholar 

  96. Wu W, LoVerde PT (2019) Nuclear hormone receptors in parasitic Platyhelminths. Mol Biochem Parasitol 233:111218. https://doi.org/10.1016/J.MOLBIOPARA.2019.111218

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Romano MC, Jiménez P, Miranda-Brito C, Valdez RA (2015) Parasites and steroid hormones: corticosteroid and sex steroid synthesis, their role in the parasite physiology and development. Front Neurosci 9:224. https://doi.org/10.3389/FNINS.2015.00224

    Article  PubMed  PubMed Central  Google Scholar 

  98. Veloz A, Reyes-Vázquez L, Patricio-Gómez JM, Romano MC (2019) Effect of mice Taenia crassiceps WFU cysticerci infection on the ovarian folliculogenesis, enzyme expression, and serum estradiol. Exp Parasitol 207:107778. https://doi.org/10.1016/J.EXPPARA.2019.107778

    Article  CAS  PubMed  Google Scholar 

  99. Morales J, Larralde C, Arteaga M, Govezensky T, Romano MC, Morali G (1996) Inhibition of sexual behavior in male mice infected with Taenia crassiceps cysticerci. J Parasitol 82:689–693. https://doi.org/10.2307/3283875

    Article  CAS  PubMed  Google Scholar 

  100. Lefèvre N, Corazza F, Valsamis J, Delbaere A, De Maertelaer V, Duchateau J, Casimir G (2019) The number of X chromosomes influences inflammatory cytokine production following toll-like receptor stimulation. Front Immunol 10:1052. https://doi.org/10.3389/FIMMU.2019.01052

    Article  PubMed  PubMed Central  Google Scholar 

  101. Rubtsova K, Marrack P, Rubtsov AV (2015) TLR7, IFNγ, and T-bet: their roles in the development of ABCs in female-biased autoimmunity. Cell Immunol 294:80–83. https://doi.org/10.1016/J.CELLIMM.2014.12.002

    Article  CAS  PubMed  Google Scholar 

  102. Spolarics Z (2007) The X-files of inflammation: cellular mosaicism of X-linked polymorphic genes and the female advantage in the host response to injury and infection. Shock 27:597–604. https://doi.org/10.1097/SHK.0B013E31802E40BD

    Article  CAS  PubMed  Google Scholar 

  103. Carrel L, Willard HF (2005) X-inactivation profile reveals extensive variability in X-linked gene expression in females. Nature 434:400–404. https://doi.org/10.1038/NATURE03479

    Article  CAS  PubMed  Google Scholar 

  104. Migeon BR (2007) Why females are mosaics, X-chromosome inactivation, and sex differences in disease. Gend Med 4:97–105. https://doi.org/10.1016/S1550-8579(07)80024-6

    Article  PubMed  Google Scholar 

  105. Mills EL, Kelly B, O’Neill LAJ (2017) Mitochondria are the powerhouses of immunity. Nat Immunol 18:488–498. https://doi.org/10.1038/NI.3704

    Article  CAS  PubMed  Google Scholar 

  106. Angajala A, Lim S, Phillips JB, Kim JH, Yates C, You Z, Tan M (2018) Diverse roles of mitochondria in immune responses: novel insights into immuno-metabolism. Front Immunol 9:1605. https://doi.org/10.3389/FIMMU.2018.01605

    Article  PubMed  PubMed Central  Google Scholar 

  107. Carnegie L, Reuter M, Fowler K, Lane N, Camus MF (2021) Mother’s curse is pervasive across a large mitonuclear Drosophila panel. Evol Lett 5:230–239. https://doi.org/10.1002/EVL3.221

    Article  PubMed  PubMed Central  Google Scholar 

  108. Vemuri R, Sylvia KE, Klein SL, Forster SC, Plebanski M, Eri R, Flanagan KL (2019) The microgenderome revealed: sex differences in bidirectional interactions between the microbiota, hormones, immunity and disease susceptibility. Semin Immunopathol 41:265–275. https://doi.org/10.1007/S00281-018-0716-7

    Article  PubMed  Google Scholar 

  109. Elderman M, Hugenholtz F, Belzer C, Boekschoten M, van Beek A, de Haan B, Savelkoul H, de Vos P, Faas M (2018) Sex and strain dependent differences in mucosal immunology and microbiota composition in mice. Biol Sex Differ 9:26. https://doi.org/10.1186/S13293-018-0186-6

    Article  PubMed  PubMed Central  Google Scholar 

  110. Zhang H, Wang Z, Li Y, Han J, Cui C, Lu C, Zhou J, Cheong L, Li Y, Sun T, Zhang D, Su X (2018) Sex-based differences in gut microbiota composition in response to tuna oil and algae oil supplementation in a d-galactose-induced aging mouse model. Front Aging Neurosci 10:187. https://doi.org/10.3389/FNAGI.2018.00187

    Article  PubMed  PubMed Central  Google Scholar 

  111. Fransen F, van Beek AA, Borghuis T, Meijer B, Hugenholtz F, van der Gaast-de JC, Savelkoul HF, de Jonge MI, Faas MM, Boekschoten MV, Smidt H, El Aidy S, de Vos P (2017) The impact of gut microbiota on gender-specific differences in immunity. Front Immunol 8:754. https://doi.org/10.3389/FIMMU.2017.00754

    Article  PubMed  PubMed Central  Google Scholar 

  112. Gause WC, Maizels RM (2016) Macrobiota—helminths as active participants and partners of the microbiota in host intestinal homeostasis. Curr Opin Microbiol 32:14–18. https://doi.org/10.1016/J.MIB.2016.04.004

    Article  PubMed  PubMed Central  Google Scholar 

  113. Reynolds LA, Smith KA, Filbey KJ, Harcus Y, Hewitson JP, Redpath SA, Valdez Y, Yebra MJ, Finlay BB, Maizels RM (2014) Commensal-pathogen interactions in the intestinal tract: lactobacilli promote infection with, and are promoted by, helminth parasites. Gut Microbes 5:522–532. https://doi.org/10.4161/GMIC.32155

    Article  PubMed  Google Scholar 

  114. Rausch S, Held J, Fischer A, Heimesaat MM, Kühl AA, Bereswill S, Hartmann S (2013) Small intestinal nematode infection of mice is associated with increased enterobacterial loads alongside the intestinal tract. PLoS One 8:e74026. https://doi.org/10.1371/JOURNAL.PONE.0074026

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Dea-Ayuela MA, Rama-Iñiguez S, Bolás-Fernandez F (2008) Enhanced susceptibility to Trichuris muris infection of B10Br mice treated with the probiotic Lactobacillus casei. Int Immunopharmacol 8:28–35. https://doi.org/10.1016/J.INTIMP.2007.10.003

    Article  CAS  PubMed  Google Scholar 

  116. Ling F, Steinel N, Weber J, Ma L, Smith C, Correa D, Zhu B, Bolnick D, Wang G (2020) The gut microbiota response to helminth infection depends on host sex and genotype. ISME J 14:1141–1153. https://doi.org/10.1038/S41396-020-0589-3

    Article  PubMed  PubMed Central  Google Scholar 

  117. Wizemann T, Pardue M (2001) Exploring the biological contributions to human health: does sex matter? National Academies Press, Washington, DC. https://doi.org/10.17226/10028

    Book  Google Scholar 

  118. Zucker I, Prendergast BJ (2020) Sex differences in pharmacokinetics predict adverse drug reactions in women. Biol Sex Differ 11:32. https://doi.org/10.1186/S13293-020-00308-5

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Fink AL, Engle K, Ursin RL, Tang WY, Klein SL (2018) Biological sex affects vaccine efficacy and protection against influenza in mice. Proc Natl Acad Sci USA 115:12477–12482. https://doi.org/10.1073/PNAS.1805268115

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Klein SL, Poland GA (2013) Personalized vaccinology: one size and dose might not fit both sexes. Vaccine 31:2599–2600. https://doi.org/10.1016/J.VACCINE.2013.02.070

    Article  PubMed  Google Scholar 

  121. Flanagan KL, Fink AL, Plebanski M, Klein SL (2017) Sex and gender differences in the outcomes of vaccination over the life course. Annu Rev Cell Dev Biol 33:577–599. https://doi.org/10.1146/ANNUREV-CELLBIO-100616-060718

    Article  CAS  PubMed  Google Scholar 

  122. Cook IF (2008) Sexual dimorphism of humoral immunity with human vaccines. Vaccine 26:3551–3555. https://doi.org/10.1016/J.VACCINE.2008.04.054

    Article  CAS  PubMed  Google Scholar 

  123. Engler RJ, Nelson MR, Klote MM, VanRaden MJ, Huang CY, Cox NJ, Klimov A, Keitel WA, Nichol KL, Carr WW, Treanor JJ, Walter Reed Health Care System Influenza Vaccine Consortium (2008) Half- vs full-dose trivalent inactivated influenza vaccine (2004–2005): age, dose, and sex effects on immune responses. Arch Intern Med 168:2405–2414. https://doi.org/10.1001/ARCHINTERNMED.2008.513

    Article  PubMed  Google Scholar 

  124. Wait LF, Dobson AP, Graham AL (2020) Do parasite infections interfere with immunisation? A review and meta-analysis. Vaccine 38:5582–5590. https://doi.org/10.1016/J.VACCINE.2020.06.064

    Article  CAS  PubMed  Google Scholar 

  125. Van Riet E, Adegnika AA, Retra K, Vieira R, Tielens AGM, Lell B, Issifou S, Hartgers FC, Rimmelzwaan GF, Kremsner PG, Yazdanbakhsh M (2007) Cellular and humoral responses to influenza in Gabonese children living in rural and semi-urban areas. J Infect Dis 196:1671–1678. https://doi.org/10.1086/522010

    Article  CAS  PubMed  Google Scholar 

  126. Hartmann W, Brunn ML, Stetter N, Gagliani N, Muscate F, Stanelle-Bertram S, Gabriel G, Breloer M (2019) Helminth infections suppress the efficacy of vaccination against seasonal influenza. Cell Rep 29:2243-2256.e4. https://doi.org/10.1016/J.CELREP.2019.10.051

    Article  CAS  PubMed  Google Scholar 

  127. vom Steeg LG, Flores-Garcia Y, Zavala F, Klein SL (2019) Irradiated sporozoite vaccination induces sex-specific immune responses and protection against malaria in mice. Vaccine 37:4468–4476. https://doi.org/10.1016/J.VACCINE.2019.06.075

    Article  Google Scholar 

  128. De SL, May S, Shah K, Slawinski M, Changrob S, Xu S, Barnes SJ, Chootong P, Ntumngia F, Adams JH (2021) Variable immunogenicity of a vivax malaria blood-stage vaccine candidate. Vaccine 39:2668–2675. https://doi.org/10.1016/J.VACCINE.2021.03.072

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Wesołowska A, Zawistowska-Deniziak A, Norbury LJ, Wilkowski P, Januszkiewicz K, Pyziel AM, Zygner W, Wędrychowicz H (2016) Immune responses in rats and sheep induced by a DNA vaccine containing the phosphoglycerate kinase gene of Fasciola hepatica and liver fluke infection. Acta Parasitol 61:212–220. https://doi.org/10.1515/AP-2016-0030

    Article  PubMed  Google Scholar 

  130. Wesołowska A, Zawistowska-Deniziak A, Norbury LJ, Wilkowski P, Pyziel AM, Zygner W, Wędrychowicz H (2018) Lymphocyte responses of rats vaccinated with cDNA encoding a phosphoglycerate kinase of Fasciola hepatica (FhPGK) and F. hepatica infection. Parasitol Int 67:85–92. https://doi.org/10.1016/J.PARINT.2017.04.002

    Article  PubMed  Google Scholar 

  131. Klein SL, Schiebinger L, Stefanick ML, Cahill L, Danska J, de Vries GJ, Kibbe MR, McCarthy MM, Mogil JS, Woodruff TK, Zucker I (2015) Opinion: sex inclusion in basic research drives discovery. Proc Natl Acad Sci USA 112:5257–5258. https://doi.org/10.1073/PNAS.1502843112

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Rich-Edwards JW, Kaiser UB, Chen GL, Manson JAE, Goldstein JM (2018) Sex and gender differences research design for basic, clinical, and population studies: essentials for investigators. Endocr Rev 39:424–439. https://doi.org/10.1210/ER.2017-00246

    Article  PubMed  PubMed Central  Google Scholar 

  133. Scott PE, Unger EF, Jenkins MR, Southworth MR, McDowell TY, Geller RJ, Elahi M, Temple RJ, Woodcock J (2018) Participation of women in clinical trials supporting FDA approval of cardiovascular drugs. J Am Coll Cardiol 71:1960–1969. https://doi.org/10.1016/J.JACC.2018.02.070

    Article  PubMed  Google Scholar 

  134. Yoon DY, Mansukhani NA, Stubbs VC, Helenowski IB, Woodruff TK, Kibbe MR (2014) Sex bias exists in basic science and translational surgical research. Surgery 156:508–516. https://doi.org/10.1016/J.SURG.2014.07.001

    Article  PubMed  Google Scholar 

  135. Beery AK (2018) Inclusion of females does not increase variability in rodent research studies. Curr Opin Behav Sci 23:143–149. https://doi.org/10.1016/j.cobeha.2018.06.016

    Article  PubMed  PubMed Central  Google Scholar 

  136. Clayton JA, Collins FS (2014) Policy: NIH to balance sex in cell and animal studies. Nature 509:282–283. https://doi.org/10.1038/509282A

    Article  PubMed  PubMed Central  Google Scholar 

  137. Schiebinger L, Klinge I (2018) Gendered innovation in health and medicine. Adv Exp Med Biol 1065:643–654. https://doi.org/10.1007/978-3-319-77932-4_39

    Article  PubMed  Google Scholar 

  138. Woitowich NC, Beery AK, Woodruff TK (2020) A 10-year follow-up study of sex inclusion in the biological sciences. Elife 9:1–8. https://doi.org/10.7554/ELIFE.56344

    Article  Google Scholar 

  139. Lubis BAA, Koesdarto S, Hestinah EP, Kusnoto K, Suwanti LT, Yunus M (2019) Prevalence of small intestine cestodes in goat at Pegirian Slaughterhouse Surabaya. J Parasite Sci 3:37. https://doi.org/10.20473/jops.v3i1.16435

    Article  Google Scholar 

  140. Ayabina DV, Clark J, Bayley H, Lamberton PH, Toor J, Hollingsworth TD (2021) Gender-related differences in prevalence, intensity and associated risk factors of Schistosoma infections in Africa: a systematic review and meta-analysis. PLoS Negl Trop Dis 15:e0009083. https://doi.org/10.1371/JOURNAL.PNTD.0009083

    Article  PubMed  PubMed Central  Google Scholar 

  141. Świsłocka M, Borkowska A, Matosiuk M, Czajkowska M, Duda N, Kowalczyk R, Ratkiewicz M (2020) Sex-biased polyparasitism in moose (Alces alces) based on molecular analysis of faecal samples. Int J Parasitol Parasites Wildl 13:171–177. https://doi.org/10.1016/J.IJPPAW.2020.10.008

    Article  PubMed  PubMed Central  Google Scholar 

  142. French AS, Zadoks RN, Skuce PJ, Mitchell G, Gordon-Gibbs DK, Taggart MA (2019) Habitat and host factors associated with liver fluke (Fasciola hepatica) diagnoses in wild red deer (Cervus elaphus) in the Scottish Highlands. Parasit Vectors 12:535. https://doi.org/10.1186/S13071-019-3782-3

    Article  PubMed  PubMed Central  Google Scholar 

  143. Jorga E, Van Damme I, Mideksa B, Gabriël S (2020) Identification of risk areas and practices for Taenia saginata taeniosis/cysticercosis in Ethiopia: a systematic review and meta-analysis. Parasites Vectors 13:375. https://doi.org/10.1186/S13071-020-04222-Y

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Lateef M, Nazir M, Zargar SA, Tariq KA (2020) Epidemiology of Taenia saginata taeniasis with emphasis on its prevalence and transmission in a Kashmiri population in India: a prospective study. Int J Infect Dis 98:401–405. https://doi.org/10.1016/J.IJID.2020.06.088

    Article  PubMed  Google Scholar 

  145. Chougar L, Harhoura K, Aissi M (2019) First isolation of Dicrocoelium dendriticum among cattle in some northern Algerian slaughterhouses. Vet World 12:1039–1045. https://doi.org/10.14202/VETWORLD.2019.1039-1045

    Article  PubMed  PubMed Central  Google Scholar 

  146. Abdelaziz AR, Khalafalla RE, Hassan AAA, Elmahallawy EK, Almuzaini AM (2019) Molecular phylogenetic analysis of Cysticercus ovis from Egypt based on MT-CO1 gene sequences. Rev Bras Parasitol Vet 28:258–265. https://doi.org/10.1590/S1984-29612019028

    Article  PubMed  Google Scholar 

  147. Khedri J, Radfar MH, Borji H, Mirzaei M (2015) Prevalence and intensity of Paramphistomum spp. in cattle from South-Eastern Iran. Iran J Parasitol 10:268–272

    PubMed  PubMed Central  Google Scholar 

  148. Kelvin EA, Carpio A, Bagiella E, Leslie D, Leon P, Andrews H, Hauser WA, Ecuadorian Neurocysticercosis Group (2013) The association of host age and gender with inflammation around neurocysticercosis cysts. Ann Trop Med Parasitol 103:487–499. https://doi.org/10.1179/000349809X12459740922291

    Article  Google Scholar 

  149. Rivero JC, Inoue Y, Murakami N, Horii Y (2002) Androgen- and estrogen-dependent sex differences in host resistance to Strongyloides venezuelensis infection in Wistar rats. J Vet Med Sci 64:457–461. https://doi.org/10.1292/JVMS.64.457

    Article  CAS  PubMed  Google Scholar 

  150. Ito EE, Egwunyenga A (2017) Soil-transmitted helminthiasis in aviara community: an observation from primary school children in Nigeria. Int Med J 24:205–208

    Google Scholar 

  151. Blancas Mosqueda M, Herrera Esparza R, Rodríguez Padilla C, Tavizón García JP, Mercado Reyes M, Badillo Almaraz V, Echavarría F, López Saucedo A, Mondragón de la Peña C (2007) Gender as a factor of susceptibility to infection in experimental hydatidosis. Rev Latinoam Microbiol 49:31–37

    PubMed  Google Scholar 

  152. Vilchez Barreto PM, Gamboa R, Santivañez S, O’Neal SE, Muro C, Lescano AG, Moyano LM, Gonzálvez G, García HH, The Cysticercosis Working Group In Perú (2017) Prevalence, age profile, and associated risk factors for Hymenolepis nana infection in a large population-based study in northern Peru. Am J Trop Med Hyg 97:583–586. https://doi.org/10.4269/ajtmh.16-0939

    Article  PubMed  PubMed Central  Google Scholar 

  153. Al-Olayan E, Elamin M, Alshehri E, Aloufi A, Alanazi Z, Almayouf M, Bakr L, Abdel-Gaber R (2020) Morphological, molecular, and pathological appraisal of Hymenolepis nana (Hymenolepididae) infecting laboratory mice (Mus musculus). Microsc Microanal 26:348–362. https://doi.org/10.1017/S1431927620000161

    Article  CAS  PubMed  Google Scholar 

  154. Radwan NA, Khalil AI, Mahi RAE (2009) Morphology and occurrence of species of Toxocara in wild mammal populations from Egypt. Comp Parasitol 76:273–282. https://doi.org/10.1654/4367.1

    Article  Google Scholar 

  155. Graham AL, Taylor MD, Le Goff L, Lamb TJ, Magennis M, Allen JE (2005) Quantitative appraisal of murine filariasis confirms host strain differences but reveals that BALB/c females are more susceptible than males to Litomosoides sigmodontis. Microbes Infect 7:612–618. https://doi.org/10.1016/J.MICINF.2004.12.019

    Article  PubMed  Google Scholar 

  156. Kofta W, Mieszczanek J, Plucienniczak G, Wȩdrychowicz H (2000) Successful DNA immunisation of rats against fasciolosis. Vaccine 18:2985–2990. https://doi.org/10.1016/S0264-410X(00)00095-5

    Article  CAS  PubMed  Google Scholar 

  157. Wedrychowicz H, Kesik M, Kaliniak M, Kozak-Cieszczyk M, Jedlina-Panasiuk L, Jaros S, Plucienniczak A (2007) Vaccine potential of inclusion bodies containing cysteine proteinase of Fasciola hepatica in calves and lambs experimentally challenged with metacercariae of the fluke. Vet Parasitol 147:77–88. https://doi.org/10.1016/J.VETPAR.2007.03.023

    Article  CAS  PubMed  Google Scholar 

  158. Jaros S, Jaros D, Wesolowska A, Zygner W, Wedrychowicz H (2010) Blocking Fasciola hepatica’s energy metabolism—a pilot study of vaccine potential of a novel gene—phosphoglycerate kinase. Vet Parasitol 172:229–237. https://doi.org/10.1016/J.VETPAR.2010.05.008

    Article  CAS  PubMed  Google Scholar 

  159. Wesołowska A, Norbury LJ, Januszkiewicz K, Jedlina L, Jaros S, Zawistowska-Deniziak A, Zygner W, Wędrychowicz H (2013) Evaluation of the immune response of male and female rats vaccinated with cDNA encoding a cysteine proteinase of Fasciola hepatica (FhPcW1). Acta Parasitol 58:198–206. https://doi.org/10.2478/S11686-013-0120-3

    Article  PubMed  Google Scholar 

  160. Wesołowska A, Basałaj K, Norbury LJ, Sielicka A, Wędrychowicz H, Zawistowska-Deniziak A (2018) Vaccination against Fasciola hepatica using cathepsin L3 and B3 proteases delivered alone or in combination. Vet Parasitol 250:15–21. https://doi.org/10.1016/J.VETPAR.2017.12.007

    Article  PubMed  Google Scholar 

  161. Wesołowska A, Basałaj K, Norbury LJ, Sielicka A, Wędrychowicz H, Zawistowska-Deniziak A (2018) Sex and vaccination: insights from female rats vaccinated with juvenile-specific proteases from Fasciola hepatica. Vet Parasitol 255:91–96. https://doi.org/10.1016/J.VETPAR.2018.04.001

    Article  PubMed  Google Scholar 

  162. Wesołowska A, Kozak Ljunggren M, Jedlina L, Basałaj K, Legocki A, Wedrychowicz H, Kesik-Brodacka M (2018) A preliminary study of a lettuce-based edible vaccine expressing the cysteine proteinase of Fasciola hepatica for fasciolosis control in livestock. Front Immunol 9:2592. https://doi.org/10.3389/FIMMU.2018.02592

    Article  PubMed  PubMed Central  Google Scholar 

  163. Josefsson E, Tarkowski A, Caristen H (1992) Anti-inflammatory properties of estrogen. I. In vivo suppression of leukocyte production in bone marrow and redistribution of peripheral blood neutrophils. Cell Immunol 142:67–78. https://doi.org/10.1016/0008-8749(92)90269-U

    Article  CAS  PubMed  Google Scholar 

  164. Jilka RL, Passeri G, Girasole G, Cooper S, Abrams J, Broxmeyer H, Manolagas SC (1995) Estrogen loss upregulates hematopoiesis in the mouse: a mediating role of IL-6. Exp Hematol 23:500–506

    CAS  PubMed  Google Scholar 

  165. Dai R, Cowan C, Heid B, Khan D, Liang Z, Pham CT, Ahmed SA (2017) Neutrophils and neutrophil serine proteases are increased in the spleens of estrogen-treated C57BL/6 mice and several strains of spontaneous lupus-prone mice. PLoS One 12:e0172105. https://doi.org/10.1371/JOURNAL.PONE.0172105

    Article  PubMed  PubMed Central  Google Scholar 

  166. Ghisletti S, Meda C, Maggi A, Vegeto E (2005) 17beta-estradiol inhibits inflammatory gene expression by controlling NF-kappaB intracellular localization. Mol Cell Biol 25:2957–2968. https://doi.org/10.1128/MCB.25.8.2957-2968.2005

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  167. Chung HH, Or YZ, Shrestha S, Loh JT, Lim CL, Ong Z, Woo ARE, Su IH, Lin VCL (2017) Estrogen reprograms the activity of neutrophils to foster protumoral microenvironment during mammary involution. Sci Rep 7:46485. https://doi.org/10.1038/SREP46485

    Article  PubMed  PubMed Central  Google Scholar 

  168. Miller AP, Feng W, Xing D, Weathington NM, Blalock JE, Chen YF, Oparil S (2004) Estrogen modulates inflammatory mediator expression and neutrophil chemotaxis in injured arteries. Circulation 110:1664–1669. https://doi.org/10.1161/01.CIR.0000142050.19488.C7

    Article  CAS  PubMed  Google Scholar 

  169. Salinas-Muñoz L, Campos-Fernández R, Mercader E, Olivera-Valle I, Fernández-Pacheco C, Matilla L, García-Bordas J, Brazil JC, Parkos CA, Asensio F, Muñoz-Fernández MA, Hidalgo A, Sánchez-Mateos P, Samaniego R, Relloso M (2018) Estrogen receptor-alpha (ESR1) governs the lower female reproductive tract vulnerability to Candida albicans. Front Immunol 9:1033. https://doi.org/10.3389/FIMMU.2018.01033

    Article  PubMed  PubMed Central  Google Scholar 

  170. Marczell I, Hrabak A, Nyiro G, Patocs A, Stark J, Dinya E, Kukor Z, Toth S, Tulassay ZS, Racz K, Bekesi G (2016) 17-β-estradiol decreases neutrophil superoxide production through Rac1. Exp Clin Endocrinol Diabetes 124:588–592. https://doi.org/10.1055/S-0042-105556

    Article  CAS  PubMed  Google Scholar 

  171. Wang J, Zhao Y, Liu C, Jiang C, Zhao C, Zhu Z (2011) Progesterone inhibits inflammatory response pathways after permanent middle cerebral artery occlusion in rats. Mol Med Rep 4:319–324. https://doi.org/10.3892/MMR.2011.418

    Article  CAS  PubMed  Google Scholar 

  172. Campbell L, Emmerson E, Williams H, Saville CR, Krust A, Chambon P, Mace KA, Hardman MJ (2014) Estrogen receptor-alpha promotes alternative macrophage activation during cutaneous repair. J Invest Dermatol 134:2447–2457. https://doi.org/10.1038/JID.2014.175

    Article  CAS  PubMed  Google Scholar 

  173. Villa A, Rizzi N, Vegeto E, Ciana P, Maggi A (2015) Estrogen accelerates the resolution of inflammation in macrophagic cells. Sci Rep 5:15224. https://doi.org/10.1038/SREP15224

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  174. Pepe G, Braga D, Renzi TA, Villa A, Bolego C, D’Avila F, Barlassina C, Maggi A, Locati M, Vegeto E (2017) Self-renewal and phenotypic conversion are the main physiological responses of macrophages to the endogenous estrogen surge. Sci Rep 7:44270. https://doi.org/10.1038/SREP44270

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Costa MC, de Barros Fernandes H, Gonçalves GKN, Santos APN, Ferreira GF, de Freitas GJC, do Carmo PHF, Hubner J, Emídio ECP, Santos JRA, Dos Santos JL, Dos Reis AM, Fagundes CT, da Silva AM, Santos DA (2020) 17-β-Estradiol increases macrophage activity through activation of the G-protein-coupled estrogen receptor and improves the response of female mice to Cryptococcus gattii. Cell Microbiol 22:e13179. https://doi.org/10.1111/CMI.13179

    Article  CAS  PubMed  Google Scholar 

  176. Capellino S, Villaggio V, Montagna P, Sulli A, Craviotto C, Cutolo M (2005) 17beta-Estradiol and testosterone influence the mRNA expression and the time course of inflammatory cytokines in activated human monocytic cell line (THP-1). Reumatismo 57:193–196. https://doi.org/10.4081/REUMATISMO.2005.193

    Article  CAS  PubMed  Google Scholar 

  177. Boje A, Moesby L, Timm M, Hansen EW (2012) Immunomodulatory effects of testosterone evaluated in all-trans retinoic acid differentiated HL-60 cells, granulocytes, and monocytes. Int Immunopharmacol 12:573–579. https://doi.org/10.1016/J.INTIMP.2012.02.008

    Article  CAS  PubMed  Google Scholar 

  178. Debelec-Butuner B, Alapinar C, Varisli L, Erbaykent-Tepedelen B, Hamid SM, Gonen-Korkmaz C, Korkmaz KS (2014) Inflammation-mediated abrogation of androgen signaling: an in vitro model of prostate cell inflammation. Mol Carcinog 53:85–97. https://doi.org/10.1002/MC.21948

    Article  CAS  PubMed  Google Scholar 

  179. Lee GT, Kim JH, Kwon SJ, Stein MN, Hong JH, Nagaya N, Billakanti S, Kim MM, Kim WJ, Kim IY (2019) Dihydrotestosterone increases cytotoxic activity of macrophages on prostate cancer cells via TRAIL. Endocrinology 160:2049–2060. https://doi.org/10.1210/EN.2019-00367

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  180. Falus A, Fehér KG, Walcz E, Brozik M, Füst G, Hidvégi T, Fehér T, Merétey K (1990) Hormonal regulation of complement biosynthesis in human cell lines–I. Androgens and gamma-interferon stimulate the biosynthesis and gene expression of C1 inhibitor in human cell lines U937 and HepG2. Mol Immunol 27:191–195. https://doi.org/10.1016/0161-5890(90)90114-F

    Article  CAS  PubMed  Google Scholar 

  181. Menzies FM, Henriquez FL, Alexander J, Roberts CW (2011) Selective inhibition and augmentation of alternative macrophage activation by progesterone. Immunology 134:281–291. https://doi.org/10.1111/J.1365-2567.2011.03488.X

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Tsai YC, Tseng JT, Wang CY, Su MT, Huang JY, Kuo PL (2017) Medroxyprogesterone acetate drives M2 macrophage differentiation toward a phenotype of decidual macrophage. Mol Cell Endocrinol 452:74–83. https://doi.org/10.1016/J.MCE.2017.05.015

    Article  CAS  PubMed  Google Scholar 

  183. Seillet C, Laffont S, Trémollières F, Rouquié N, Ribot C, Arnal JF, Douin-Echinard V, Gourdy P, Guéry JC (2012) The TLR-mediated response of plasmacytoid dendritic cells is positively regulated by estradiol in vivo through cell-intrinsic estrogen receptor α signaling. Blood 119:454–464. https://doi.org/10.1182/BLOOD-2011-08-371831

    Article  CAS  PubMed  Google Scholar 

  184. Laffont S, Rouquié N, Azar P, Seillet C, Plumas J, Aspord C, Guéry JC (2014) X-Chromosome complement and estrogen receptor signaling independently contribute to the enhanced TLR7-mediated IFN-α production of plasmacytoid dendritic cells from women. J Immunol 193:5444–5452. https://doi.org/10.4049/JIMMUNOL.1303400

    Article  CAS  PubMed  Google Scholar 

  185. Mackern-Oberti JP, Jara EL, Riedel CA, Kalergis AM (2017) Hormonal modulation of dendritic cells differentiation, maturation and function: implications for the initiation and progress of systemic autoimmunity. Arch Immunol Ther Exp (Warsz) 65:123–136. https://doi.org/10.1007/S00005-016-0418-6

    Article  CAS  Google Scholar 

  186. Bupp MRG, Jorgensen TN (2018) Androgen-induced immunosuppression. Front Immunol 9:794. https://doi.org/10.3389/FIMMU.2018.00794

    Article  Google Scholar 

  187. Corrales JJ, Almeida M, Burgo R, Mories MT, Miralles JM, Orfao A (2006) Androgen-replacement therapy depresses the ex vivo production of inflammatory cytokines by circulating antigen-presenting cells in aging type-2 diabetic men with partial androgen deficiency. J Endocrinol 189:595–604. https://doi.org/10.1677/JOE.1.06779

    Article  CAS  PubMed  Google Scholar 

  188. Arck P, Hansen PJ, Mulac Jericevic B, Piccinni MP, Szekeres-Bartho J (2007) Progesterone during pregnancy: endocrine-immune cross talk in mammalian species and the role of stress. Am J Reprod Immunol 58:268–279. https://doi.org/10.1111/J.1600-0897.2007.00512.X

    Article  CAS  PubMed  Google Scholar 

  189. Bouman A, Jan Heineman M, Faas MM (2005) Sex hormones and the immune response in humans. Hum Reprod Update 11:411–423. https://doi.org/10.1093/HUMUPD/DMI008

    Article  CAS  PubMed  Google Scholar 

  190. Curran EM, Berghaus LJ, Vernetti NJ, Saporita AJ, Lubahn DB, Estes DM (2001) Natural killer cells express estrogen receptor-alpha and estrogen receptor-beta and can respond to estrogen via a non-estrogen receptor-alpha-mediated pathway. Cell Immunol 214:12–20. https://doi.org/10.1006/CIMM.2002.1886

    Article  CAS  PubMed  Google Scholar 

  191. Jiang X, Orr BA, Kranz DM, Shapiro DJ (2006) Estrogen induction of the granzyme B inhibitor, proteinase inhibitor 9, protects cells against apoptosis mediated by cytotoxic T lymphocytes and natural killer cells. Endocrinology 147:1419–1426. https://doi.org/10.1210/EN.2005-0996

    Article  CAS  PubMed  Google Scholar 

  192. Laskarin G, Tokmadzić VS, Strbo N, Bogović T, Szekeres-Bartho J, Randić L, Podack ER, Rukavina D (2002) Progesterone induced blocking factor (PIBF) mediates progesterone induced suppression of decidual lymphocyte cytotoxicity. Am J Reprod Immunol 48:201–209. https://doi.org/10.1034/J.1600-0897.2002.01133.X

    Article  PubMed  Google Scholar 

  193. Bonds RS, Midoro-Horiuti T (2013) Estrogen effects in allergy and asthma. Curr Opin Allergy Clin Immunol 13:92–99. https://doi.org/10.1097/ACI.0B013E32835A6DD6

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Kamis AB, Ibrahim JB (1989) Effects of testosterone on blood leukocytes in plasmodium berghei-infected mice. Parasitol Res 75:611–613. https://doi.org/10.1007/BF00930957

    Article  CAS  PubMed  Google Scholar 

  195. Hirokuni N, Yoichiro H, Koichiro F (1992) Effect of testosterone on the eosinophil response of C57BL/6 mice to infection with Brugia pahangi. Immunopharmacology 23:75–79. https://doi.org/10.1016/0162-3109(92)90030-G

    Article  Google Scholar 

  196. Becerra-Díaz M, Strickland AB, Keselman A, Heller NM (2018) Androgen and androgen receptor as enhancers of M2 macrophage polarization in allergic lung inflammation. J Immunol 201:2923–2933. https://doi.org/10.4049/JIMMUNOL.1800352

    Article  PubMed  Google Scholar 

  197. Karpuzoglu-Sahin E, Hissong BD, Ansar Ahmed S (2001) Interferon-gamma levels are upregulated by 17-beta-estradiol and diethylstilbestrol. J Reprod Immunol 52:113–127. https://doi.org/10.1016/S0165-0378(01)00117-6

    Article  CAS  PubMed  Google Scholar 

  198. Karpuzoglu E, Phillips RA, Gogal RM, Ansar Ahmed S (2007) IFN-gamma-inducing transcription factor, T-bet is upregulated by estrogen in murine splenocytes: role of IL-27 but not IL-12. Mol Immunol 44:1808–1814. https://doi.org/10.1016/J.MOLIMM.2006.08.005

    Article  CAS  PubMed  Google Scholar 

  199. Butts CL, Shukair SA, Duncan KM, Bowers E, Horn C, Belyavskaya E, Tonelli L, Sternberg EM (2007) Progesterone inhibits mature rat dendritic cells in a receptor-mediated fashion. Int Immunol 19:287–296. https://doi.org/10.1093/INTIMM/DXL145

    Article  CAS  PubMed  Google Scholar 

  200. Sakazaki F, Ueno H, Nakamuro K (2008) 17beta-Estradiol enhances expression of inflammatory cytokines and inducible nitric oxide synthase in mouse contact hypersensitivity. Int Immunopharmacol 8:654–660. https://doi.org/10.1016/J.INTIMP.2008.01.007

    Article  CAS  PubMed  Google Scholar 

  201. Lambert KC, Curran EM, Judy BM, Milligan GN, Lubahn DB, Estes DM (2005) Estrogen receptor alpha (ERalpha) deficiency in macrophages results in increased stimulation of CD4+ T cells while 17beta-estradiol acts through ERalpha to increase IL-4 and GATA-3 expression in CD4+ T cells independent of antigen presentation. J Immunol 175:5716–5723. https://doi.org/10.4049/JIMMUNOL.175.9.5716

    Article  CAS  PubMed  Google Scholar 

  202. Matalka KZ (2003) The effect of estradiol, but not progesterone, on the production of cytokines in stimulated whole blood, is concentration-dependent. Neuro Endocrinol Lett 24:185–191

    CAS  PubMed  Google Scholar 

  203. Dalal M, Kim S, Voskuhl RR (1997) Testosterone therapy ameliorates experimental autoimmune encephalomyelitis and induces a T helper 2 bias in the autoantigen-specific T lymphocyte response. J Immunol 159:3–6

    CAS  PubMed  Google Scholar 

  204. Konermann A, Winter J, Novak N, Allam JP, Jäger A (2013) Verification of IL-17A and IL-17F in oral tissues and modulation of their expression pattern by steroid hormones. Cell Immunol 285:133–140. https://doi.org/10.1016/J.CELLIMM.2013.10.004

    Article  CAS  PubMed  Google Scholar 

  205. Massa MG, David C, Jörg S, Berg J, Gisevius B, Hirschberg S, Linker RA, Gold R, Haghikia A (2017) Testosterone differentially affects T cells and neurons in murine and human models of neuroinflammation and neurodegeneration. Am J Pathol 187:1613–1622. https://doi.org/10.1016/J.AJPATH.2017.03.006

    Article  CAS  PubMed  Google Scholar 

  206. Jia T, Anandhan A, Massilamany C, Rajasekaran RA, Franco R, Reddy J (2015) Association of autophagy in the cell death mediated by dihydrotestosterone in autoreactive T cells independent of antigenic stimulation. J Neuroimmune Pharmacol 10:620–634. https://doi.org/10.1007/S11481-015-9633-X

    Article  PubMed  PubMed Central  Google Scholar 

  207. Lajko A, Meggyes M, Polgar B, Szereday L (2018) The immunological effect of Galectin-9/TIM-3 pathway after low dose Mifepristone treatment in mice at 14.5 day of pregnancy. PLoS One 13:e0194870. https://doi.org/10.1371/JOURNAL.PONE.0194870

    Article  PubMed  PubMed Central  Google Scholar 

  208. Tai P, Wang J, Jin H, Song X, Yan J, Kang Y, Zhao L, An X, Du X, Chen X, Wang S, Xia G, Wang B (2008) Induction of regulatory T cells by physiological level estrogen. J Cell Physiol 214:456–464. https://doi.org/10.1002/JCP.21221

    Article  CAS  PubMed  Google Scholar 

  209. Walecki M, Eisel F, Klug J, Baal N, Paradowska-Dogan A, Wahle E, Hackstein H, Meinhardt A, Fijak M (2015) Androgen receptor modulates Foxp3 expression in CD4+CD25+Foxp3+ regulatory T-cells. Mol Biol Cell 26:2845–2857. https://doi.org/10.1091/MBC.E14-08-1323

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  210. Marguti I, Yamamoto GL, da Costa TB, Rizzo LV, de Moraes LV (2009) Expansion of CD4+ CD25+ Foxp3+ T cells by bone marrow-derived dendritic cells. Immunology 127:50–61. https://doi.org/10.1111/J.1365-2567.2008.02927.X

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  211. Khan D, Ansar Ahmed S (2016) The immune system is a natural target for estrogen action: opposing effects of estrogen in two prototypical autoimmune diseases. Front Immunol 6:635. https://doi.org/10.3389/FIMMU.2015.00635

    Article  PubMed  PubMed Central  Google Scholar 

  212. Hill L, Jeganathan V, Chinnasamy P, Grimaldi C, Diamond B (2011) Differential roles of estrogen receptors α and β in control of B-cell maturation and selection. Mol Med 17:211–220. https://doi.org/10.2119/MOLMED.2010.00172

    Article  CAS  PubMed  Google Scholar 

  213. Jones BG, Penkert RR, Xu B, Fan Y, Neale G, Gearhart PJ, Hurwitz JL (2016) Binding of estrogen receptors to switch sites and regulatory elements in the immunoglobulin heavy chain locus of activated B cells suggests a direct influence of estrogen on antibody expression. Mol Immunol 77:97–102. https://doi.org/10.1016/J.MOLIMM.2016.07.015

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  214. Medina KL, Kincade PW (1994) Pregnancy-related steroids are potential negative regulators of B lymphopoiesis. Proc Natl Acad Sci USA 91:5382–5386. https://doi.org/10.1073/PNAS.91.12.5382

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  215. Zhang L, Chang KK, Li MQ, Li DJ, Yao XY (2014) Mouse endometrial stromal cells and progesterone inhibit the activation and regulate the differentiation and antibody secretion of mouse B cells. Int J Clin Exp Pathol 7:123–133

    PubMed  Google Scholar 

Download references

Funding

This work was supported by a National Science Centre Grant (project number: 2019/35/B/NZ6/04002).

Author information

Authors and Affiliations

Authors

Contributions

AW performed the literature review and wrote the manuscript. The author read and approved the final manuscript.

Corresponding author

Correspondence to Agnieszka Wesołowska.

Ethics declarations

Competing interests

The author declares no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1.

Effects of sex steroids on immune cells. File contains a table that summarizes the effects of androgens, oestrogens, and progesterone on different immune cells.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wesołowska, A. Sex—the most underappreciated variable in research: insights from helminth-infected hosts. Vet Res 53, 94 (2022). https://doi.org/10.1186/s13567-022-01103-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13567-022-01103-3

Keywords