Skip to main content

Sex-specific influence on cardiac structural remodeling and therapy in cardiovascular disease

Abstract

Background

Cardiovascular diseases (CVDs) culminating into heart failure (HF) are major causes of death in men and women. Prevalence and manifestation, however, differ between sexes, since men mainly present with coronary artery disease (CAD) and myocardial infarction (MI), and post-menopausal women predominantly present with hypertension. These discrepancies are probably influenced by underlying genetic and molecular differences in structural remodeling pathways involved in hypertrophy, inflammation, fibrosis, and apoptosis. In general, men mainly develop eccentric forms, while women develop concentric forms of hypertrophy. Besides that, women show less inflammation, fibrosis, and apoptosis upon HF. This seems to emerge, at least partially, from the fact that the underlying pathways might be modulated by estrogen, which changes after menopause due to declining of the estrogen levels.

Conclusion

In this review, sex-dependent alterations in adverse cardiac remodeling are discussed for various CVDs. Moreover, potential therapeutic options, like estrogen treatment, are reviewed.

Background

Although sex-differences in heart failure (HF) and underlying cardiovascular diseases (CVDs) is not an unexplored topic, sex-based therapies are still remarkably scarce. Independent of the injury, CVDs lead to adverse cardiac remodeling, which can eventually culminate into HF. HF is classified based on selected symptoms, as described for example by the Task Force of the European Society of Cardiology, and the left ventricular ejection fraction (LVEF) separating into HF with preserved LVEF (> 50%, HFpEF), HF with moderate LVEF (40–50%, HFmEF), and HF with reduced LVEF (< 40%, HFrEF) [1]. Epidemiology and pathophysiology of CVDs and therefore HF often differ between sexes [2, 3]. Until now, sex-based discrepancies in the essential molecular pathways of CVDs such as coronary artery disease (CAD), myocardial infarction (MI), hypertension, valvular heart disease, chronic obstructive pulmonary disease (COPD), and different forms of cardiomyopathy (namely, arrhythmogenic (ACM), hypertrophic (HCM), and dilated (DCM)) are incompletely understood, aggravating the development of proper treatment for both sexes.

Many of these sex-differences are caused by the presence or absence of sex hormones, but besides sex hormones of course also other factors can influence adverse cardiac remodeling in a sex-specific manner, such as sex chromosomes and epigenetics. It has been shown that several genes related to, e.g., hypertension and adverse cardiac remodeling processes, such as macrophage activation, apoptosis, and lipid metabolisms are located on the Y chromosome [4]. Moreover, lack of the X chromosome can lead to various diseases as is further elaborated on in the discussion, and incomplete X chromosome inactivation can lead to disease progression as reviewed in, e.g., Regitz-Zagrosek and Kararigas [5]. However, the role of the chromosomes and epigenetics, as reviewed by, e.g., Hartman et al. [6], is beyond the scope of this review.

Also metabolism, pregnancy, life-style, and environmental factors play an important contributing role on sex-differences in the most common CVDs. These influences have been reviewed extensively before [5, 7, 8]. Therefore, our review specifically aims at summarizing sex differences in the molecular remodeling mechanisms, which are particularly involved in hypertrophy, inflammation, apoptosis, and fibrosis.

Sex differences in patients with CVD

When generalized, prevalence of CVDs and overall mortality are independent of sex [3]. However, epidemiology and pathophysiology of CVDs often differ as has been extensively reviewed in literature and is briefly summarize below and in Table 1.

Table 1 Sex differences in cardiovascular structural remodeling

CAD and MI are more prevalent in men than in women in all age groups (43% in men versus 27% in women in the CONFIRM Long-Term Registry) and women acquire those diseases about 10 years later in life, though overall mortality has shown to be equal specifically for CAD and MI (about 50% in Writing Group 2016) [3, 9, 10]. In the CONFIRM study, women who presented with CAD were older (62.4 years versus 58.9 years in men) and suffered more from hypertension (60% versus 52% in men) [9]. After MI, women often develop HFpEF (37% versus 23% in men), whereas men mainly present with HFrEF [11]. Prevalence of hypertension keeps increasing and is the most treated risk factor for CVD and HF, leading to mainly HFpEF in women and HFrEF in men [12]. Pre-menopausal women have a lower incidence of hypertension than age-matched men (e.g., 7.3% versus 15.8% men, between 30 and 44 years of age) [13]. This difference vanishes during menopause until the age of 65, when more women present with hypertension than men. This fact is mainly thought to be caused by the influence of estrogens in females, but might also partly be explained by a genetic component as is suggested through the identification of a hypertension associated region on the Y chromosome in animals [4, 14, 15]. Reports on valvular heart diseases are contradictory, indicating that more women suffer from mitral valve disease (male:female ratio, 3:1), and depending on the different studies, either more men or women present with aortic valve disease (male:female ratio, 3:1 or 0.76:1) [3, 14].

Pulmonary hypertension causes right ventricular (RV) cardiac remodeling, with pulmonary stenosis and COPD being the two main contributors. In contrast to systemic hypertension, pulmonary hypertension is more prevalent in younger women (< 65 years) than in men (male:female ratio, 1:1.8 in the UK), but pregnancy in younger women highly contributes to these numbers. However, at older ages, the prevalence equalizes [16]. In both sexes, the disease often culminates into HFpEF, and men present with a higher mortality, despite the lower prevalence [1]. In the non-smoking group, about 80% of COPD patients are female [17].

Prevalence of the different forms of cardiomyopathies, such as DCM, HCM, and ACM is higher, average age of diagnosis is lower, and outcome is worse in men than in women (male:female ratio of DCM 1.5:1, of HCM 3:2, and of ACM 3:1) [18]. Interestingly, inherited forms of these cardiomyopathies, e.g., due to mutations in troponin T (for DCM), Plakophilin-2 (for ACM) and in sarcomeric proteins (for HCM), also show a skewed incidence, suggesting that on top of the genetic predisposition, additional factors are decisive for the phenotypical expression of the disease [18].

Regardless of the cardiovascular injury, women tend to develop HFpEF, while most men develop HFrEF (e.g., 73.4% males developed HFrEF and 48% developed HFpEF in the PREVEND study [19]) as depicted in Fig. 1. Whether women developing HFrEF or men developing HFpEF show the same characteristics of cardiac remodeling as their counterparts is currently still under investigation (indicated in with a question mark).

Fig. 1
figure 1

Generalized development of sex differences in structural cardiac remodeling and possible heart failure (HF) outcomes. Men (blue lines) and women (red lines) often show different structural remodeling after an initial injury (e.g., hypertension and MI.) leading to acute inflammation and different disease outcomes. In general, women tend to develop HF later than men. The older, the more patients suffer from cardiovascular diseases; however, in women this seems to be clearly related to the onset of their menopause. Thickness of lines mirrors the general number of patients suffering from the HF type at that age

Sex differences in molecular pathways of adverse cardiac remodeling

CVDs trigger adverse cardiac remodeling, which is manifested by hypertrophy, inflammation, fibrosis, and apoptosis, eventually leading to HF. Many of these processes are often sex-specific (see Table 1) and influenced by sex-hormone levels of estrogen, testosterone, or progesterone [20]. Obviously, men and women are on different hormonal schedules, and although the hormonal shifts during a woman’s life are more frequent and dramatic than in a men’s life, it is thought that the effects of sex-hormones, in particular estrogen, during the pre-menopausal period are able to delay the appearance of CVDs in women [15]. This is supported by the observation that women with irregular hormone cycles and an early menopause have an elevated risk of HF [20].

In the following sections, we will discuss the mentioned mechanisms and processes triggering adverse cardiac remodeling in the light of sex-specific differences.

Hypertrophy

The development and manifestation of cardiac hypertrophy differs between the sexes. Generally, in women, LV remodeling due to MI or aortic stenosis/hypertension is oriented towards concentric hypertrophy leading to diastolic dysfunction (Fig. 1 pink lines). In men, eccentric hypertrophy with chamber dilatation is the most common, eventually leading to systolic dysfunction (Fig. 1 blue lines) [1, 21, 22]. Next to that, the amount of viable cardiomyocytes decreases in the healthy aging male heart more extensively than in the female heart, making it at baseline already more susceptible to stress and hypertrophy [23]. As mentioned, this effect in females is thought to be, at least in part, due to estrogens and estrogen receptor (ER) signaling, which is especially mediated by the estrogen receptor beta (ER-β) [24]. Studies in animal models support this. Mice lacking the ER-β develop hypertrophy faster, and females with aortic stenosis show a more extensive upregulation of ER-β, while ER-α is upregulated to the same extent in both sexes [25, 26]. Estrogen-treated ovariectomized (OVX) female mice show a slower development of LV hypertrophy compared to untreated OVX animals [27]. In patients, estrogen is associated with a better RV systolic function, and androgens, such as testosterone, with an increased RV mass and volume, as is seen during hypertrophy [28]. Moreover, estrogen is known to modulate natriuretic peptides, e.g., pro-brain-type natriuretic peptide (BNP) and activates angiogenesis, which supports the hypertrophic heart in its elevated oxygen demand [29]. Estrogen has also been shown to temper hypertrophic remodeling through CaMKII inhibition, which adds to the preserved EF in pre-menopausal women [30].

Pre-menopausal women have a lower incidence of hypertension than age-matched men as mentioned earlier. This is in line with observations that estrogen can modulate β-adrenergic signaling and may cause vasodilation by increasing endothelial nitric oxide synthase (eNOS) [31]. LV pressure overload leads in general to activation of the neuro-hormonal axis and to an increase of the sympathetic drive. In young men, peripheral resistance is proportional to sympathetic drive, though not in young women. Post-menopausal peripheral resistance increases with sympathetic drive [32]. This is probably due to the decrease of estrogen, since baroreflex sensitivity is modulated by estrogens [33]. Because of that, post-menopausal women have a higher prevalence of hypertension and pulmonary hypertension. In addition, men with aortic stenosis show a more robust upregulation of genes related to the renin–angiotensin–aldosterone system (RAAS), whereas women are able to repress the gene encoding for angiotensinogen [34]. Transverse aortic constriction (TAC), mimicking aortic stenosis in mice and rats, shows in the majority of studies that female animals develop HF later than males, accompanied with a less dilated phenotype [35].

Although estrogen is known for its vasodilator properties, pulmonary hypertension is more prevalent in younger women, and the role of estrogens and testosterone in pulmonary hypertension remains controversial. After hypoxia-induced pulmonary artery hypertension in mice with overexpressed serotonin transporters, OVX leads to a reduced RV hypertrophy, whereas estrogen therapy (ET) re-established the hypertension phenotype [36, 37]. In contrast, in rats with pulmonary hypertension, experimentally induced by, e.g., monocrotaline or bleomycin, estrogen has been reported to be beneficial against RV hypertrophy, inflammation, and fibrosis, and females show a less severe phenotype than males [38]. Besides that, these different outcomes might be related to differences in the type of rodent and the method used to induce hypertension; in numerous studies, female animals show less pulmonary hypertension than male animals as reviewed in Mair et al. [39]. However, the same investigators suggest that men are probably more protected from pulmonary hypertension due to the fact that the vasodilatory effects of testosterone are more important in the pulmonary vasculature than that of estrogens [39].

HCM presents by definition with concentric and DCM with eccentric hypertrophy. Both disorders are more prevalent in men as discussed before. Also in mice, females seem to develop less DCM and HCM than males, and later in life upon end-stage HF induced by chronic LV pressure overload [18]. ACM leads to predominantly RV dilatation in humans and mice (e.g., upon Plakophilin-2 knockout) [40], but also biventricular phenotypes are known [41]. In ACM, dilatation may be caused by increased stretch of the cardiac walls and be more visible in the right ventricle, since this wall is significantly thinner [42]. In men, this effect is often more pronounced, which might be due to a higher intrinsic sympathetic drive and activity in young men and male children [43].

Although studies in animals highlight potential mechanisms of cardiovascular remodeling mediated by sex-hormones, these models often make use of one sex-hormone at a time and are therefore not optimal representations of the fine balance between different sex-hormones under physiological conditions. Extrapolations to the human condition should therefore be considered with care.

Inflammation

Inflammatory pathways contribute to the progression of adverse remodeling and are activated upon cardiac injury. As mentioned before, the Y chromosome contains several genes involved in, e.g., macrophage activation [4]. Higher numbers of neutrophils and macrophages in males lead to an increased extent of inflammation, and together with a reportedly higher activity of matrix metalloproteinases (MMPs), this can cause premature breakdown of collagen and therefore cardiac rupture and slowed healing after injury or during HF [44].

The most common inflammatory pathways triggering secretion of cytokines and chemokines and subsequently the attraction of inflammatory cells like dendritic cells and macrophages are depicted in Fig. 2 and discussed in the following paragraphs.

Fig. 2
figure 2

Molecular signaling of inflammatory pathways involved in structural remodeling. Inflammatory pathways activated upon cardiac injury include the IFN-γ/JAK-STAT pathway (purple), the TLR pathway (green; either in an endosome (left pathway) or on the cell membrane (right pathway)), the TNF-α/MEKK pathway (red), and the TGF-β/BMP/SMAD pathway (blue), triggering secretion of cytokines and chemokines and leading to, e.g., apoptosis and cell proliferation. Crosstalk between the pathways is possible (dashed arrow) and collaboration with the Wnt pathway (light gray) and CaMKII (dark gray) promoting cardiac inflammation. Sex differences in these pathways are discussed in the text, such as the inhibitory effects of estrogen on CaMKII, BNP, and NFκB

The JAK-STAT pathway

Binding of a ligand to the interferon (IFN) receptor leads to activation of the Janus kinase-signal transducer and activator of transcription (JAK-STAT) pathway, causing translocation of a STAT dimer to the nucleus (Fig. 2, purple pathway). A study showed that in women undergoing LV adverse remodeling due to pressure overload, genes of the JAK-STAT pathway are significantly downregulated [34]. Mechanical strain upon hemodynamic stress (such as in TAC animal models) induces production of the cytokine interleukin-6 (IL-6), activating the JAK-STAT pathway, and also the toll-like receptor (TLR) pathway (Fig. 2, green pathway) [45]. Female mice lacking the ER-β have significantly increased activation of inflammatory signals, such as IL-6, and presence of ER-β seems to prevent cardiac inflammation and fibrosis in mice upon TAC [25, 46].

The toll-like receptor pathway

The inflammatory TLR pathway activation is associated with adverse cardiac remodeling, arrhythmias, and HF [47]. Binding of a ligand to a TLR (either in an endosome (Fig. 2, left green pathway) or on the cell membrane (Fig. 2, green pathway)) leads to either translocation of IRF3 or AP1 to the nucleus, or degradation of the inhibitor of NFκB (IκB) and consequently transport of NFκB to the nucleus. In women, the TLR pathways are to a lesser extent activated than in men, and male patients show increased expression of various inflammatory proteins upon LV remodeling (e.g., chemokine receptor-1, IL-33, and TLR7) [34]. In animals with experimentally induced hypertension and MI, TLR expression is increased in males, while knockout of TLRs (e.g., TLR4 and TLR2) improves cardiac function more in males than in females after MI [48,49,50,51]. Interestingly, ER-α can act as a transcription factor, which can inhibit expression of NFκB [52].

Crosstalk between different remodeling mechanisms is noticed since hypertrophy can interact with inflammatory signaling, where CaMKII (which is prominently activated during adverse remodeling) can directly phosphorylate components of the TLR pathway, thereby promoting cardiac inflammation (Fig. 2, black arrow) [53]. As mentioned above, estrogens can interact with CaMKII and members of the TLR pathway, therefore playing an important part in this interaction. In the future, therapeutic interference with these pathways, as exampled by TLR blockers, should therefore take the sex of the patient and his/her hormone levels into account.

The Wnt pathway

During inflammation, the Wnt pathway (Fig. 2, gray pathway) is activated and known to regulate T cell development, dendritic cell, and leukocyte maturation [54]. Activation of the Wnt pathway leads to degradation of the GSK3-β complex resulting in distortion of catenins from the membrane and translocation to the nucleus. Men with aortic stenosis show a larger upregulation of the Wnt pathway than women [34]. Crosstalk between the Wnt/β-catenin and the TLR pathway has been reported, where depletion of β-catenin leads to increased levels of TLR4 [54]. In ACM patients, mutation of γ-catenin (also known as Plakoglobin) or β-catenin causes weakening of intercellular junctions, probably leading to detachment of cardiac cells. This causes cell death and necrosis, which has been associated with inflammation and eventually fibrosis formation, although sex differences in this regard have not been investigated [55]. Not only interactions between GSK3 and ER-α, but also androgen receptors have been reviewed in detail [56], and this should be taken into account when investigating for example GSK3-β blockers. Furthermore, Wnt signaling is involved in a plethora of (developmental) processes, making it a difficult target for therapeutic interventions.

The TNF pathway

Binding to the tumor necrosis factor (TNF) receptor activates TRADD (Fig. 2, red pathway), which in turn either activates FADD and the apoptotic caspase pathway or TRAF2, which leads to cross activation of the TLR (green pathway) and TGF-β and bone morphogenic protein (BMP) pathways (Fig. 2, blue pathway). Also upon MI, the apoptotic pathway is activated by caspase-11 that in turn can activate caspase-1. The latter is able to induce interleukin (e.g., IL-1α, IL-6) and TNF-α production, and their secretion causes activation of more inflammatory cells and pathways, such as the TLR and MEKK pathway [57]. Estrogen has been shown to inhibit the synthesis of TNF-α, thereby hampering the activation of inflammatory cells and fibroblasts [58]. This suggests that in women, the TNF pathway is less active compared to men and therefore the development/progression of CVDs might be slowed. Repressing the TNF pathway therapeutically might therefore have a greater benefit in men and post-menopausal women than pre-menopausal women.

The TGF-β and BMP pathway

Invading inflammatory cells, such as macrophages, express the TGF-β receptor, which is able to mediate the inflammatory response [59]. Activation of the TGF-β or BMP receptor leads to SMAD complex formation, activation of Ras and the mitogen-activated protein kinase pathway, or activation of PI3K (Fig. 2, blue pathways). This eventually results in epithelial to mesenchymal transition or activation of the mTOR pathway causing changes in cell cycle and proliferation. Women show lower expression of TGF-β and a lower inflammatory response [34] probably since ER-β binding by estrogen on fibroblasts has been shown to block TGF-β and AngII stimulation [46]. In addition, TGF-β is highly pro-fibrotic and macrophages play a role in transition of fibroblasts to myofibroblasts; the latter being more pro-fibrotic in their activity [59]. This illustrates the tight connection between inflammatory and fibrotic responses in cardiac remodeling, logically leading to less of both in women, or at least a delay in response when compared to men.

Concluding, all cardiac inflammatory pathways can be modulated by estrogen and therefore, sex differences should be taken into account when conducting research on and/or testing therapeutic interventions on CVDs.

Fibrosis and apoptosis

During all forms of CVDs and HF, and in both sexes, cardiac fibroblasts secrete excessive amounts of collagen into the extracellular matrix. While collagen genes are upregulated in all kinds of human CVDs and rodent disease models, Metalloproteinases (MMPs) are often downregulated or unchanged, and Tissue Inhibitors of MMPs (TIMPs) are either upregulated or also unchanged, as thoroughly reviewed in Fan et al [60]. Besides, although different CVDs result in different patterns of fibrosis, its formation and deposition appears later in the process of remodeling in women as compared to men [34].

Apoptosis is the programmed cell death following, e.g., cardiac injury, and in general, female failing hearts show lower rates of apoptosis and necrosis than males. This could, on the one hand, be due to the presence of apoptosis-related genes on the Y chromosome, or on the other hand, although not completely understood, due to the role of estrogen and ERs [4].

After MI in mouse models, mostly compact fibrosis is found due to massive apoptosis and necrosis in the infarcted area, which is more extensive and results in more collagen content in male mice compared to females [61]. This is further confirmed in humans, where apoptosis after MI is more prevalent in men compared to women [62]. After pressure overload, female rodents present with lower levels of apoptotic proteins, such as caspase 3 and 9, and ER-β signaling attenuates apoptosis and seems to prevent cardiac fibrosis upon chronic pressure overload and in HF [25, 63, 64]. In OVX rats, estrogen therapy is able to reduce collagen content and MMP/TIMP balance upon LV pressure overload and apoptotic protein levels [64, 65]. In addition, as already explained, females show lower levels of TGF-β and angiotensinogen, both involved in synthesis and deposition of fibrosis [34]. Male patients undergoing aortic valve surgery, on the other hand, display increased levels of collagens I and III, and MMP expression as compared to female patients [66]. In male HFpEF patients, also expression of TIMP-2 appeared to be increased, as compared to females [34]. Pulmonary hypertension leads to RV fibrosis and apoptosis in both sexes [67]. Reports on the role of sex-hormones in pulmonary hypertension are contradictory: since prevalence is higher in females, one would suggest a protective role for, e.g., testosterone. However, castrated rats with pulmonary hypertension showed the opposite since decreased RV fibrosis was observed while administration of testosterone lead to an increase in collagen content [68].

Women with DCM have less fibrosis and apoptosis, and show a more reversible type of HF in clinics [69]. In DCM patients with lamin A/C mutations, nuclear accumulation of androgens such as testosterone has been found, and castrated mice carrying the same mutations showed improved cardiac function and less cell death and fibrosis [70]. Structural sex-based differences in the clinical manifestation of ACM, such as the disease-characteristic fibro-fatty replacement and apoptosis, have not been reported [71], probably since detection is predominantly based on MRI which allows only the detection of larger areas of fibrosis. However, in a recent study, testosterone worsened and estradiol improved the degree of cardiac apoptosis and lipogenesis in an ACM model of induced pluripotent stem cells [72]. Moreover, lipid metabolism seems to be associated with loci on the sex chromosomes, possibly leading to sex-specific differences in fibro-fatty replacement [4].

Apoptosis and fibrosis in the myocardium cause structural abnormalities and inhomogeneity of the syncytium, which is an important factor that affects efficiency of contractility and predisposes to arrhythmogenesis [73]. Therefore, delayed deposition of fibrosis and apoptosis in female hearts as compared to males might postpone the appearance of such cardiac dysfunctions and should be targeted in future research in a sex-specific manner.

Sex differences in therapeutic interventions

Pharmacological interventions for CVDs and HF include prescription of statins, neuro-hormonal blockage by, e.g., angiotensin-converting enzyme (ACE) inhibitors, angiotensin receptor blockers and β-blockers, and instrumentation via, e.g., an implantable cardiac defibrillator (ICD) in order to prevent sudden death caused by arrhythmic events. In general, except for diuretics, most drugs are more often prescribed in men than in women, even after adjusting for age and mortality. Often absorption, distribution, metabolism, and clearance differ between sexes and sex-specific differences have been reviewed in, for instance Tamargo et al. [74], and are shortly summarized in Table 2. Unfortunately, current guidelines are not considering these differences and therapeutic interferences like neuro-hormonal blockage are to a lesser extent beneficial in patients with HFpEF and therefore, in general, less effective in women than in men [3]. Moreover, women have a lower distribution volume for medications, such as β-blockers, related to differences in body dimensions and fat distribution, and often show a lower clearance rate [75]. Prescription of drugs varies between sexes, where statins, for example are more likely prescribed to men than women (64.8% in men versus 57.6% in women). Furthermore, of all ICD implantations, only 17–26% are performed in women [76,77,78]. The latter most likely can be explained by the fact that guidelines for ICD therapy (LVEF < 35%) are based on the results of clinical trials, where men are overrepresented.

Table 2 Sex differences in therapeutic options for CVDs

Pharmacological treatment for pulmonary hypertension mainly includes calcium channel blockers, whose mode of action can be increased in women due to higher bioavailability [16, 74].

Therapy for ACM aims at preventing arrhythmias and reducing the rate of SCD with, e.g., exercise restrictions and ICD therapy [41]. Since ICD therapy in ACM is often used as primary prophylaxis, sex-biased outcomes are not that prominent.

Since many differences exist between the response and therapies, and therefore their efficiency between women and men, it is surprising that there are no sex-specific guidelines to treat CVDs and HF. Most of the therapeutic options are driven by the manifestation of CVDs in men, while they are also the primary cause of death in women [1]. Future sex-specific therapeutic options should therefore focus on the underlying differences in cardiac remodeling, but one should also keep in mind that often the primary injury seems to be already different between sexes. This is, e.g., the case in CAD, where women show more often microvascular and men macrovascular dysfunction [10]. Furthermore, biomarker reference intervals for these diseases have not historically been established for men and women separately, which may partially contribute to the gaps in underlying diagnosis.

This together illustrates that sometimes a different treatment-regimen will be needed for men and women.

Estrogen therapy (ET) has been investigated as a potential treatment for CVDs in post-menopausal women as extensively reviewed in Miller and Harman [79]. In rodents with chronic pressure overload, ET seems to increase EF and reduce fibrosis formation [80]. However, controversial results in clinical trials using ET show either improvement of contractile performance upon, e.g., hypertension or an actual increase in the risk of stroke, MI, and pulmonary embolism [81]. The effect might be dependent on the type of estrogen administered (e.g., differently dosed 17-β estradiol (1 or 2 mg), estradiol valerate, conjugated equine estrogen, or combined therapies), the way of administration (orally or non-orally), and the timing/frequency. Besides that, positive effects of ET are often only seen in clinical trials, if estrogen is administered within 1 year after menopause suggesting that it is not possible to reverse some permanent alterations during disease progression (e.g., fibrosis or accumulation of atherosclerotic plaques) [82].

Furthermore, in “extreme patients,” such as individuals with congenital Turner syndrome, the effect of lack of estrogen can be highlighted. Women with Turner syndrome lack one X chromosome partially or fully and often suffer from premature ovarian failure leading to estrogen deficiency or even depletion [83]. These women often present with hypertension are at risk for various cardiac diseases and have a 3-fold increased mortality. Hormone replacement therapy and ET is a standard therapy that is already initiated early in life. ET in the affected women does not only normalize, e.g., uterine size, but was also shown to improve blood pressure and, at a high dose, increased the cardiac intima media thickness as intensively reviewed in Cintron et al. [83]

In line with that, women with congenital Kallmann syndrome are characterized by hypogonadotropic hypogonadism and anosmia. Moreover, they often suffer from congenital heart diseases. ET is also used in these cases already early in life to induce secondary sex characteristics, but less is known about its potential role in cardioprotection [84].

Conclusions and future perspectives

Although sex-specific discrepancies are seen in almost all CVDs, appropriate diagnosis and consequent therapies are not well established. This is not only partly due to severe underrepresentation of women in clinical trials (young women were even banned from clinical trials by the FDA from 1977 to 1993) [85], but also caused by the lack of knowledge on molecular disparities between sexes with regard to cardiac remodeling pathways for hypertrophy, inflammation, fibrosis, and apoptosis. In this review, we summarized the knowledge on sex-dependent differences in these pathways and the involvement of sex-hormones, especially estrogen. Overall, women show less fibrosis, hypertrophy, and inflammation upon CVDs than men, often preserving their EF. This leads to an onset of a most often different phenotype of HF later in life, but on the other hand prevents women from receiving the appropriate timely treatment, since the absence of a proper HF diagnosis means that no treatment will be provided. This increases the risk of sudden cardiac death in females. Therefore, also women with HFpEF should be treated or at least be monitored to reduce mortality in this group.

Moreover, women seem to seek less cardiac health care, treatment (such as an ICD) is often less well perceived and induces more psychological discomfort (depression and stress) than in men, and men are more often heart donors and receivers [14, 35]. Partially, this could rely on the fact that different cardiac symptoms are experienced by women and men, which could hamper the recognition in women that they actually are experiencing, e.g., an MI. These facts demonstrate that sex-specific differences in the underlying mechanisms of structural remodeling indeed lead to differences in CVDs and health care and should be more thoroughly investigated to advance and specify treatment.

Abbreviations

ACE:

Angiotensin-converting enzyme

ACM:

Arrhythmogenic cardiomyopathy

BNP:

Pro-brain-type natriuretic peptide

CAD:

Coronary artery disease

CaMK II:

Calmodulin kinase 2

COPD:

Chronic obstructive pulmonary disease

CVD:

Cardiovascular disease

DCM:

Dilated cardiomyopathy

ECM:

Extra cellular matrix

eNOS:

Endothelial nitric oxide synthase

ER:

Estrogen receptor

ET:

Estrogen therapy

ET-1:

Endothelin-1

HCM:

Hypertrophic cardiomyopathy

HF:

Heart failure

HFmEF:

Heart failure with moderate LVEF

HFpEF:

Heart failure with preserved LVEF

HFrEF:

Heart failure with reduced LVEF

ICD:

Implantable cardiac defibrillator

IFN-γ:

Interferon-γ

IL:

Interleukin

JAK-STAT:

Janus kinase/signal transducer and activator of transcription

LV:

Left ventricle

LVEF:

Left ventricular ejection fraction

MEKK:

Mitogen-activated protein kinase kinase

MI:

Myocardial infarction

MMP:

Matrix metalloproteinase

MRI:

Magnetic resonance imaging

NF-κB:

Nuclear factor kappa-light-chain-enhancer of activated B cells

OVX:

Ovariectomized

PLN:

Phospholamban

RAAS:

Renin–angiotensin–aldosterone system

RV:

Right ventricle

SCD:

Sudden cardiac death

TAC:

Transverse aortic constriction

TGF-β:

Transforming growth factor-β

TIMP:

Tissue inhibitor of MMP

TLR:

Toll-like receptor

References

  1. Ponikowski P, Voors AA, Anker SD, Bueno H, Cleland JGF, Coats AJS, et al. 2016 ESC guidelines for the diagnosis and treatment of acute and chronic heart failure. Eur Heart J. 2016;37:2129–200.

    Article  PubMed  Google Scholar 

  2. Bleumink GS, Knetsch AM, Sturkenboom MCJM, Straus SMJM, Hofman A, Deckers JW, et al. Quantifying the heart failure epidemic: prevalence, incidence rate, lifetime risk and prognosis of heart failure—the Rotterdam study. Eur Heart J. 2004;25(18):1614–9.

    Article  PubMed  Google Scholar 

  3. Writing Group Members, Mozaffarian D, Benjamin EJ, Go AS, Arnett DK, Blaha MJ, et al. Heart disease and stroke statistics-2016 update: a report from the American Heart Association. Circulation. 2016;133(4):e38–360.

    Google Scholar 

  4. Winham SJ, de Andrade M, Miller VM. Genetics of cardiovascular disease: importance of sex and ethnicity. Atherosclerosis. 2014;241(1):219–28.

    Article  CAS  Google Scholar 

  5. Regitz-Zagrosek V, Kararigas G. Mechanistic pathways of sex differences in cardiovascular disease. Physiol Rev. 2017;97(1):1–37.

    Article  PubMed  Google Scholar 

  6. Hartman RJG, Huisman SE, Den Ruijter HM. Sex differences in cardiovascular epigenetics—a systematic review. Biol Sex Differ. 2018;9:19.

    Article  PubMed  PubMed Central  Google Scholar 

  7. Chella Krishnan K, Mehrabian M, Lusis AJ. Sex differences in metabolism and cardiometabolic disorders. Curr Opin Lipidol. 2018;29(5):404–10.

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Appelman Y, van Rijn BB, ten Haaf ME, Boersma E, Peters SAE. Sex differences in cardiovascular risk factors and disease prevention. Atherosclerosis. 2014;241(1):211–8.

    Article  CAS  Google Scholar 

  9. Schulman-Marcus J, Hartaigh BÓ, Gransar H, Lin F, Valenti V, Cho I, et al. Sex-specific associations between coronary artery plaque extent and risk of major adverse cardiovascular events the CONFIRM long-term registry. JACC Cardiovasc Imaging. 2016;9(4):364–72.

    Article  PubMed  PubMed Central  Google Scholar 

  10. Maas AH, Appelman YEA. Gender differences in coronary heart disease. Neth Heart J. 2010;18(12):598–603.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Hellermann JP, Jacobsen SJ, Reeder GS, Lopez-Jimenez F, Weston SA, Roger VL. Heart failure after myocardial infarction: prevalence of preserved left ventricular systolic function in the community. Am Heart J. 2003;145(4):742–8.

    Article  PubMed  Google Scholar 

  12. Lam CSP, Donal E, Kraigher-Krainer E, Vasan RS. Epidemiology and clinical course of heart failure with preserved ejection fraction. Eur Heart Fail. 2011;13(1):18–28.

    Article  Google Scholar 

  13. Neuhauser H, Thamm M, Ellert U. Blood pressure in Germany 2008-2011: results of the German Health Interview and Examination Survey for Adults (DEGS1). Bundesgesundheitsblatt Gesundheitsforschung Gesundheitsschutz. 2013;56(5–6):795–801.

    Article  CAS  PubMed  Google Scholar 

  14. Regitz-Zagrosek V, Oertelt-Prigione S, Prescott E, Franconi F, Gerdts E, Foryst-Ludwig A, et al. Gender in cardiovascular diseases: impact on clinical manifestations, management, and outcomes. Eur Heart J. 2016;37(1):24–34.

    Article  CAS  PubMed  Google Scholar 

  15. Mendelsohn ME, Karas RH. The protective effects of estrogen on the cardiovascular system. N Engl J Med. 1999;340(23):1801–11.

    Article  CAS  PubMed  Google Scholar 

  16. Opitz C, Rosenkranz S, Ghofrani HA, Grünig E, Klose H, Olschewski H, et al. ESC guidelines 2015 pulmonary hypertension: diagnosis and treatment. Dtsch Med Wochenschr. 2016;141(24):1764–9.

    Article  PubMed  Google Scholar 

  17. Aryal S, Diaz-Guzman E, Mannino DM. COPD and gender differences: an update. Transl Res. 2013;162(4):208–18.

    Article  PubMed  Google Scholar 

  18. Meyer S, van der Meer P, van Tintelen JP, van den Berg MP. Sex differences in cardiomyopathies. Eur J Heart Fail. 2014;16(3):238–47.

    Article  PubMed  Google Scholar 

  19. Brouwers FP, De Boer RA, Van Der Harst P, Voors AA, Gansevoort RT, Bakker SJ, et al. Incidence and epidemiology of new onset heart failure with preserved vs. reduced ejection fraction in a community-based cohort: 11-year follow-up of PREVEND. Eur Heart J. 2013;34(19):1424–31.

    Article  CAS  PubMed  Google Scholar 

  20. Appiah D, Schreiner PJ, Demerath EW, Loehr LR, Chang PP, Folsom AR. Association of age at menopause with incident heart failure: a prospective cohort study and meta-analysis. J Am Heart Assoc. 2016;5(8):e003769.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  21. Crabbe DL, Dipla K, Ambati S, Zafeiridis A, Gaughan JP, Houser SR, et al. Gender differences in post-infarction hypertrophy in end-stage failing hearts. J Am Coll Cardiol. 2003;41(2):300–6.

    Article  PubMed  Google Scholar 

  22. Aurigenvna GP, Gaasch WH. Gender differences in older patients with pressure-overload hypertrophy of the left ventricle. Cardiol. 1995;86(4):310–7.

    Article  Google Scholar 

  23. Olivetti G, Melissari M, Capasso JM, Anversa P. Cardiomyopathy of the aging human heart. Myocyte loss and reactive cellular hypertrophy. Circ Res. 1991;68(6):1560–8.

    Article  CAS  PubMed  Google Scholar 

  24. Skavdahl M, Steenbergen C, Clark J, Myers P, Demianenko T, Mao L, et al. Estrogen receptor-β mediates male-female differences in the development of pressure overload hypertrophy. Am J Physiol Circ Physiol. 2005;288(2):469–76.

    Article  CAS  Google Scholar 

  25. Kararigas G, Fliegner D, Gustafsson J-A, Regitz-Zagrosek V. Role of the estrogen/estrogen-receptor-beta axis in the genomic response to pressure overload-induced hypertrophy. Physiol Genomics. 2011;43(8):438–46.

    Article  CAS  PubMed  Google Scholar 

  26. Mahmoodzadeh S, Eder S, Nordmeyer J, Ehler E, Huber O, Martus P, et al. Estrogen receptor alpha up-regulation and redistribution in human heart failure. FASEB J. 2006;20(7):926–34.

    Article  CAS  PubMed  Google Scholar 

  27. Van Eickels M, Grohé C, Cleutjens JPM, Janssen BJ, Wellens HJJ, Doevendans PA. 17β-estradiol attenuates the development of pressure-overload hypertrophy. Circulation. 2001;104(12):1419–23.

    Article  PubMed  Google Scholar 

  28. Ventetuolo CE, Ouyang P, Bluemke DA, Tandri H, Barr RG, Bagiella E, et al. Sex hormones are associated with right ventricular structure and function: the MESA-right ventricle study. Am J Respir Crit Care Med. 2011;183(5):659–67.

    Article  PubMed  Google Scholar 

  29. Kuroski De Bold ML. Estrogen, natriuretic peptides and the renin-angiotensin system. Cardiovasc Res. 1999;41(3):524–31.

    Article  CAS  PubMed  Google Scholar 

  30. Pedram A, Razandi M, Narayanan R, Dalton JT, McKinsey TA, Levin ER. Estrogen regulates histone deacetylases to prevent cardiac hypertrophy. Mol Biol Cell. 2013;24(24):3805–18.

    Article  PubMed  PubMed Central  Google Scholar 

  31. Mendelsohn ME. Genomic and nongenomic effects of estrogen in the vasculature. Am J Cardiol. 2002;90(1A):3F–6F.

    Article  CAS  PubMed  Google Scholar 

  32. Shoemaker JK, Hogeman CS, Khan M, Kimmerly DS, Sinoway LI. Gender affects sympathetic and hemodynamic response to postural stress. Am J Physiol Heart Circ Physiol. 2001;281(5):2028–35.

    Article  Google Scholar 

  33. Barnes JN, Matzek LJ, Charkoudian N, Joyner MJ, Curry TB, Hart EC. Association of cardiac baroreflex sensitivity with blood pressure transients: influence of sex and menopausal status. Front Physiol. 2012;3:187.

    Article  PubMed  PubMed Central  Google Scholar 

  34. Kararigas G, Dworatzek E, Petrov G, Summer H, Schulze TM, Baczko I, et al. Sex-dependent regulation of fibrosis and inflammation in human left ventricular remodelling under pressure overload. Eur J Heart Fail. 2014;16(11):1160–7.

    Article  CAS  PubMed  Google Scholar 

  35. Regitz-Zagrosek V, Seeland U. Sex and gender differences in myocardial hypertrophy and heart failure. Wien Med Wochenschr. 2011;161(5–6):109–16.

    Article  PubMed  Google Scholar 

  36. White K, Dempsie Y, Nilsen M, Wright AF, Loughlin L, MacLean MR. The serotonin transporter, gender, and 17β oestradiol in the development of pulmonary arterial hypertension. Cardiovasc Res. 2011;90(2):373–82.

    Article  CAS  PubMed  Google Scholar 

  37. English KM, Jones RD, Jones TH, Morice AH, Channer KS. Gender differences in the vasomotor effects of different steroid hormones in rat pulmonary and coronary arteries. Horm Metab Res. 2001;33(11):645–52.

    Article  CAS  PubMed  Google Scholar 

  38. Tofovic SP, Zhang X, Jackson EK, Zhu H, Petrusevska G. 2-methoxyestradiol attenuates bleomycin-induced pulmonary hypertension and fibrosis in estrogen-deficient rats. Vasc Pharmacol. 2009;51(2–3):190–7.

    Article  CAS  Google Scholar 

  39. Mair KM, Johansen AKZ, Wright AF, Wallace E, Maclean MR. Pulmonary arterial hypertension: basis of sex differences in incidence and treatment response. Br J Pharmacol. 2014;171(3):567–79.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Cerrone M, Noorman M, Lin X, Chkourko H, Liang FX, Van Der Nagel R, et al. Sodium current deficit and arrhythmogenesis in a murine model of plakophilin-2 haploinsufficiency. Cardiovasc Res. 2012;95(4):460–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Corrado D, Wichter T, Link MS, Hauer R, Marchlinski F, Anastasakis A, et al. Treatment of arrhythmogenic right ventricular cardiomyopathy/dysplasia: an international task force consensus statement. Circulation. 2015;132:441–53.

    Article  PubMed  PubMed Central  Google Scholar 

  42. Calkins H. Arrhythmogenic right ventricular dysplasia/cardiomyopathy—three decades of Progress. Circ J. 2015;79(5):901–13.

    Article  PubMed  Google Scholar 

  43. Bauce B, Frigo G, Marcus FI, Basso C, Rampazzo A, Maddalena F, et al. Comparison of clinical features of arrhythmogenic right ventricular cardiomyopathy in men versus women. Am J Cardiol. 2008;102(9):1252–7.

    Article  PubMed  Google Scholar 

  44. Cavasin MA, Tao Z, Menon S, Yang XP. Gender differences in cardiac function during early remodeling after acute myocardial infarction in mice. Life Sci. 2004;75(18):2181–92.

    Article  CAS  PubMed  Google Scholar 

  45. Nagai T, Anzai T, Kaneko H, Mano Y, Anzai A, Maekawa Y, et al. C-reactive protein overexpression exacerbates pressure overload-induced cardiac remodeling through enhanced inflammatory response. Hypertension. 2011;57(2):208–15.

    Article  CAS  PubMed  Google Scholar 

  46. Pedram A, Razandi M, O’Mahony F, Lubahn D, Levin ER. Estrogen receptor-β prevents cardiac fibrosis. Mol Endocrinol. 2010;24(11):2152–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Avlas O, Bragg A, Fuks A, Nicholson JD, Farkash A, Porat E, et al. TLR4 expression is associated with left ventricular dysfunction in patients undergoing coronary artery bypass surgery. PLoS One. 2015;10(6):e0120175.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  48. Li J, Ao L, Zhai Y, Clevel JC, Fullerton DA, Meng X. Gender disparity in the role of TLR2 in post-ischemic myocardial inflammation and injury. Int J Clin Exp Med. 2015;8(7):10537–47.

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Liu L, Wang Y, Cao ZY, Wang MM, Liu XM, Gao T, et al. Up-regulated TLR4 in cardiomyocytes exacerbates heart failure after long-term myocardial infarction. J Cell Mol Med. 2015;19(12):2728–40.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Eiler R, Schmaderer C, Rusai K, Kühne L, Sollinger D, Lahmer T, et al. Hypertension augments cardiac toll-like receptor 4 expression and activity. Hypertens Res. 2011;34(5):551–8.

    Article  CAS  Google Scholar 

  51. Ha T, Li Y, Hua F, Ma J, Gao X, Kelley J, et al. Reduced cardiac hypertrophy in toll-like receptor 4-deficient mice following pressure overload. Cardiovasc Res. 2005;68(2):224–34.

    Article  CAS  PubMed  Google Scholar 

  52. Kalaitzidis D, Gilmore TD. Transcription factor cross-talk: the estrogen receptor and NF-κB. Trends Endocrin Metab. 2005;16(2):46–52.

    Article  CAS  Google Scholar 

  53. Liu X, Yao M, Li N, Wang C, Zheng Y, Cao X. CaMKII promotes TLR-triggered proinflammatory cytokine and type I interferon production by directly binding and activating TAK1 and IRF3 in macrophages. Blood. 2008;112(13):4961–70.

    Article  CAS  PubMed  Google Scholar 

  54. Ma B, Hottiger MO. Crosstalk between wnt/β-catenin and NF-κB signaling pathway during inflammation. Front Immunol. 2016;7:378.

    PubMed  PubMed Central  Google Scholar 

  55. Thiene G, Nava A, Corrado D, Rossi L, Pennelli N. Right ventricular cardiomyopathy and sudden death in young people. N Engl J Med. 1988;318(3):129–33. 6.

    Article  CAS  PubMed  Google Scholar 

  56. Beurel E, Grieco SF, Jope RS. Glycogen synthase kinase-3 (GSK3): Regulation, actions, and diseases. Pharmacol Ther. 2015. https://doi.org/10.1016/j.pharmthera.2014.11.016.

  57. Gottlieb R. ICE-ing the heart. Circ Res. 2005;96(10):1036–8.

    Article  CAS  PubMed  Google Scholar 

  58. McLarty JL, Li J, Levick SP, Janicki JS. Estrogen modulates the influence of cardiac inflammatory cells on function of cardiac fibroblasts. J Inflamm Res. 2013;6(1):99–108.

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Dobaczewski M, Chen W, Frangogiannis NG. Transforming growth factor (TGF)-β signaling in cardiac remodeling. JMCC. 2011;51(4):600–6.

    CAS  Google Scholar 

  60. Fan D, Takawale A, Lee J, Kassiri Z. Cardiac fibroblasts, fibrosis and extracellular matrix remodeling in heart disease. Fibrogenesis Tissue Repair. 2012;5(1):15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Weber KT, Sun Y, Díez J. Fibrosis. A living tissue and the infarcted heart. J Am Coll Cardiol. 2008;52(24):2029–31.

    Article  PubMed  Google Scholar 

  62. Biondi-Zaccai GGL, Abbate A, Bussani R, Camilot D, Giorgio FD, Marino MPD, et al. Reduced post-infarction myocardial apoptosis in women: a clue to their different clinical course? Heart. 2005;91(1):99–101.

    Article  Google Scholar 

  63. Pedram A, Razandi M, Narayanan R, Levin ER. Estrogen receptor beta signals to inhibition of cardiac fibrosis. Mol Cell Endocrinol. 2016;434:57–68.

    Article  CAS  PubMed  Google Scholar 

  64. Dent MR, Tappia PS, Dhalla NS. Gender differences in apoptotic signaling in heart failure due to volume overload. Apoptosis. 2010;15(4):499–510.

    Article  PubMed  Google Scholar 

  65. Voloshenyuk TG, Gardner JD. Estrogen improves TIMP-MMP balance and collagen distribution in volume-overloaded hearts of ovariectomized females. Am J Physiol Integr Comp Physiol. 2010;299(2):683–93.

    Article  CAS  Google Scholar 

  66. Petrov G, Regitz-Zagrosek V, Lehmkuhl E, Krabatsch T, Dunkel A, Dandel M, et al. Regression of myocardial hypertrophy after aortic valve replacement: Faster in women? Circulation. 2010;122(11 SUPPL. 1):S23–8.

    Article  PubMed  Google Scholar 

  67. Shehata ML, Lossnitzer D, Skrok J, Boyce D, Lechtzin N, Mathai SC, et al. Myocardial delayed enhancement in pulmonary hypertension: pulmonary hemodynamics, right ventricular function, and remodeling. Am J Roentgenol. 2011;196(1):87–94.

    Article  Google Scholar 

  68. Hemnes AR, Maynard KB, Champion HC, Gleaves L, Penner N, West J, et al. Testosterone negatively regulates right ventricular load stress responses in mice. Pulm Circ. 2012;2(3):352–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Fairweather DL, Cooper LT, Blauwet LA. Sex and gender differences in myocarditis and dilated cardiomyopathy. Curr Probl Cardiol. 2013;38(1):7–46.

    Article  PubMed  PubMed Central  Google Scholar 

  70. Arimura T, Onoue K, Takahashi-Tanaka Y, Ishikawa T, Kuwahara M, Setou M, et al. Nuclear accumulation of androgen receptor in gender difference of dilated cardiomyopathy due to Lamin a/C mutations. Cardiovasc Res. 2013;99(3):382–94.

    Article  CAS  PubMed  Google Scholar 

  71. Lin CY, Chung FP, Lin YJ, Chang SL, Lo LW, Hu YF, et al. Gender differences in patients with arrhythmogenic right ventricular dysplasia/cardiomyopathy: clinical manifestations, electrophysiological properties, substrate characteristics, and prognosis of radiofrequency catheter ablation. Int J Cardiol. 2017;227:930–7.

    Article  PubMed  Google Scholar 

  72. Akdis D, Saguner AM, Shah K, Wei C, Medeiros-Domingo A, Von Eckardstein A, et al. Sex hormones affect outcome in arrhythmogenic right ventricular cardiomyopathy/dysplasia: from a stem cell derived cardiomyocyte-based model to clinical biomarkers of disease outcome. Eur Heart J. 2017;38(19):1498–508.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Tamarappoo BK, John BT, Reinier K, Teodorescu C, Uy-Evanado A, Gunson K, et al. Vulnerable myocardial interstitium in patients with isolated left ventricular hypertrophy and sudden cardiac death: a postmortem histological evaluation. J Am Heart Assoc. 2012;1(3):e001511.

    Article  PubMed  PubMed Central  Google Scholar 

  74. Tamargo J, Rosano G, Walther T, Duarte J, Niessner A, Kaski JC, et al. Gender differences in the effects of cardiovascular drugs. Eur Heart J Cardiovasc Pharmacother. 2017;3(3):163–82.

    Article  CAS  PubMed  Google Scholar 

  75. Soldin OP, Chung SH, Mattison DR. Sex differences in drug disposition. J Biomed Biotechnol. 2011. https://doi.org/10.1155/2011/187103.

  76. Amit G, Suleiman M, Konstantino Y, Luria D, Kazatsker M, Chetboun I, et al. Sex differences in implantable cardioverter-defibrillator implantation indications and outcomes: lessons from the nationwide Israeli-ICD registry. Europace. 2014;16(8):1175–80.

    Article  PubMed  Google Scholar 

  77. Varma N, Mittal S, Prillinger JB, Snell J, Dalal N, Piccini JP. Survival in women versus men following implantation of pacemakers, defibrillators, and cardiac resynchronization therapy devices in a large, nationwide cohort. J Am Heart Assoc. 2017;6(5):e005031.

    Article  PubMed  PubMed Central  Google Scholar 

  78. Virani SS, Woodard LD, Ramsey DJ, Urech TH, Akeroyd JM, Shah T, et al. Gender disparities in evidence-based statin therapy in patients with cardiovascular disease. Am J Cardiol. 2015;115(1):21–6.

    Article  PubMed  Google Scholar 

  79. Miller VM, Harman SM. An update on hormone therapy in postmenopausal women: mini-review for the basic scientist. Am J Physiol Heart Circ Physiol. 2017;313(5):H1013–21.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  80. Iorga A, Li J, Sharma S, Umar S, Bopassa JC, Nadadur RD, et al. Rescue of pressure overload-induced heart failure by estrogen therapy. J Am Heart Assoc. 2016;5(1):1–12.

    Article  Google Scholar 

  81. Boardman HMP, Hartley L, Eisinga A, Main C, Roqué i Figuls M, Bonfill Cosp X, et al. Hormone therapy for preventing cardiovascular disease in post-menopausal women. Cochrane Database Syst Rev. 2015;10(3):CD002228.

    Google Scholar 

  82. Salpeter SR, Walsh JME, Greyber E, Salpeter EE. Brief rport: Cronary heart disease events associated with hormone therapy in younger and older women—a meta-analysis. J Gen Intern Med. 2006;21(4):363–6.

    Article  PubMed  PubMed Central  Google Scholar 

  83. Cintron D, Rodriguez-Gutierrez R, Serrano V, Latortue-Albino P, Erwin PJ, Murad MH. Effect of estrogen replacement therapy on bone and cardiovascular outcomes in women with turner syndrome: a systematic review and meta-analysis. Endocrine. 2017;55(2):366–75.

    Article  CAS  PubMed  Google Scholar 

  84. Meczekalski B, Podfigurna-Stopa A, Smolarczyk R, Katulski K, Genazzani AR. Kallmann syndrome in women: from genes to diagnosis and treatment. Gynecol Endocrinol. 2013;29(4):296–300.

    Article  PubMed  Google Scholar 

  85. Liu KA, Dipietro Mager NA. Women’s involvement in clinical trials: historical perspective and future implications. Pharm Pract. 2016;14(1):708.

    Article  Google Scholar 

  86. Olsson KM, Channick R. Pregnancy in pulmonary arterial hypertension. Eur Resp Rev. 2016;25:431–7.

    Article  Google Scholar 

  87. Patten RD. Models of gender differences in cardiovascular disease. Drug Discov Today Dis Models. 2007;4(4):227–32.

    Article  PubMed  PubMed Central  Google Scholar 

  88. Austin ED, Lahm T, West J, Tofovic SP, Johansen AK, Maclean MR, et al. Gender, sex hormones and pulmonary hypertension. Pulm Circ. 2013;3(2):294–314.

    Article  PubMed  PubMed Central  Google Scholar 

  89. Sheppard R, Bedi M, Kubota T, Semigran MJ, Dec W, Holubkov R, et al. Myocardial expression of fas and recovery of left ventricular function in patients with recent-onset cardiomyopathy. J Am Coll Cardiol. 2005;46(6):1036–42.

    Article  CAS  PubMed  Google Scholar 

  90. Campian ME, Verberne HJ, Hardziyenka M, De Groot EAA, Van Moerkerken AF, Van Eck-Smit BLF, et al. Assessment of inflammation in patients with arrhythmogenic right ventricular cardiomyopathy/dysplasia. Eur J Nucl Med Mol Imaging. 2010;37(11):2079–85.

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

None.

Funding

This work was supported by the Netherlands CardioVascular Research Initiative: the Dutch Heart Foundation, Dutch Federation of University Medical Centers, the Netherlands Organization for Health Research and Development and the Royal Netherlands Academy of Sciences (CVON-PREDICT 2012-10).

Availability of data and materials

Not applicable.

Author information

Authors and Affiliations

Authors

Contributions

ELK made substantial contributions to conception and design, and drafted the manuscript. MRR was involved in conception and design, and drafting the manuscript or revising it critically for important intellectual content. MAV was involved in revising the manuscript critically for important intellectual content. TABvV made substantial contributions to conception and design of the manuscript and was involved in drafting the manuscript or revising it critically for important intellectual content. All authors gave final approval of the version to be published and agreed to be accountable for all aspects of the work in ensuring that questions related to the accuracy or integrity of any part of the work are appropriately investigated and resolved.

Corresponding author

Correspondence to Elise L. Kessler.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

All authors give consent for publication for all content.

Competing interests

The authors declare that they have no competing interests.

Publisher’s Note

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

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. 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.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kessler, E.L., Rivaud, M.R., Vos, M.A. et al. Sex-specific influence on cardiac structural remodeling and therapy in cardiovascular disease. Biol Sex Differ 10, 7 (2019). https://doi.org/10.1186/s13293-019-0223-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13293-019-0223-0

Keywords