Keywords

1 Introduction

To date, the term ‘cardiac remodelling’ (CR) was firstly coined by Hockman and Buckey on myocardial infarction (MI) to study replacement of myocardial injury with scar tissue [1]. Later, Janice Pfeffer applied the name CR to illustrate the progressive dilatation of the left ventricle (LV) on vivo studies [2]. Nonetheless, Pfeffer and Braunwald utilized the term CR for morphological changes caused by MI, especially for LV remodelling [3]. In this regard, an international forum published in 2000 a consensus on CR, that defined “CR as a group of molecular, cellular and interstitial changes that clinically manifest as changes in size, shape and function of the heart resulting from cardiac injury” [4]. In biology, the term remodelling characterizes adjustments that cause reorganizing of initial structures [5]. Even if CR was used initially to describe the geometric and structural modifications caused by MI [3, 6], CR is actually applied to a large variety of cardiac conditions. Esentially, Swynghedauw classified etiology of CR as being (1) acquired diseases (postmyocardial infarction, hypertensive cardiopathy, valve and congenital disease, myocarditis, and Chagas disease); (2) genetics (inherited cardiomyopathies, familial hypertrophic cardiomyopathy, dilated cardiomyopathy, Marfan disease, hemochromatose, transgenic models, transgenic models of cardiac hypertrophy, transgenic models of cardiac failure); and (3) miscellaneous causes (aging; heart rate; use of catecholamines, thyroxine, or growth hormone; salt, mineralo- and glucocorticoid; diabetes mellitus; B6 vitamin deficiency; atrophy due to heterotopic transplantation and hypertrophy due to homeotopic transplantation (?)) [5, 7, 8]. Further, CR is separated into structural (hypertrophy and fibrosis) and electrical remodelling . Shortly, any type of stress induces cardiomyocytes (CMs) to become hypertrophic with altered electrical function, while cardiac fibroblasts (CFs) transform in ‘activated’ myofibroblasts (MyoFb), which further multiply and boost extracellular matrix (ECM) tissue with fibrosis [9].

During international forum from 2000 [4], two types of CR were established: (1) physiological (adaptive) remodelling and (2) pathological remodelling. Further, Dorn et al. defined CR as being ‘adaptive or maladaptive’ [10]. It should be noted, that Hill and Olson stated that heart can respond to environmental stimuli by increase of myocardial mass or atrophy starting with a “least 100%” [11]. More important is other mechanisms than remodelling also can alter the evolution of heart disease, even in the absence of remodelling process. To reiterate, CR can be a physiologic or pathologic condition [4]. Physiologic CR is a physiological alteration in size and function of the heart due to physiologic stimuli such as exercise (“athlete’s heart”) and pregnancy. In addition, pathologic CR occurs with pressure overload conditions (e.g., aortic stenosis, hypertension), with volume overload conditions (e.g., valvular regurgitation), with cardiac injury or coronary artery disease (CAD) , and with inflammatory myocardial disease (e.g., myocarditis), or idiopathic dilated cardiomyopathy [4]. Equally, physiologic CR may lead to pathologic remodelling [12].

Constrictive pericardial disease, selected forms of congenital heart diseases (CHD) , inflow obstruction, primary myocardial disease, and pressure or volume overload are each well-described causes of right ventricular (RV) remodelling , RV systolic dysfunction, and cor pulmonale [13]. Emerging evidence suggests that RV dysfunction is the mainly marker of poor prognosis in pulmonary hypertension (PH) [14, 15].

For simplicity, the first adaptive reaction of the RV to pressure overload is hypertrophy. If untreated, the RV dilates to compensate increased RV preload and to maintain stroke volume according to the Frank-Starling principle. When further increase in RV end-diastolic filling volume do not balance progressive RV contractile dysfunction, clinically evident RV failure ensues. In advanced stages, RV dilation may also impair LV diastolic filling kinetics that contributes further to global pump dysfunction and, consequently, to the congestive heart failure (CHF) syndrome [16].

It should be restated that the RV and the LV don’t have same embryologic origins. The RV stems from the secondary/anterior heart tube and the LV from early/primary heart tube [17]. Accordingly, RV formation is specifically controlled by several genes, including Hand2 and Tbx20 [18]. This different embryologic origin of RV is associated with cellular divergence that controls the duration of early development to different LV and RV cardiomyocytes, and go on with distinct cell signalling and Ca2+ handling pathways for both chambers, altogether suggestive of certain essential differentiation at the cellular level for both ventricles [19]. For the foetal period, the RV propels blood into the pulmonary circulation, placenta and into the inferior body. Further, during the switch from the foetal circulation to the postnatal circulation with the reduction of pulmonary vascular resistance (PVR), the RV develops into a thin-walled, heavily trabeculated chamber pushing a cardiac output (CO) same to the LV but with lesser energy cost [20]. Normal crescent–shaped RV with thinner walls is a low-pressure chamber that faces the low impedance of pulmonary circulation. Thus, although the RV is a low-resistance and low-capacitance pump, the LV is an high-resistance and high-pressure pump [20]. Additionally, the RV has a different metabolism and morphology in comparison with LV [20]. RV cardiomyoytes are disposed longitudinally and demonstrate faster twitch velocities than the radially oriented LV cardiomyocytes. As a result, because of these anatomical and physiological differences, both ventricles present various reactions to disease forms. According to current evidences, it seems that RV hypertrophy (RVH), RV remodelling and RV failure (RVF) can develop at the same time instead of progression development (Fig. 4.1) [21].

Fig. 4.1
figure 1

(a) An overview of changes associated with RV pressure overload. Key triggers of RV pressure overload include pulmonary hypertension, RV outflow tract obstruction or RV being the systemic ventricle. RV pressure overload induces RVH that, through remodelling, leads to RV failure. It is of note, however, that RV failure is a continuous process and may begin as the time of hypertrophy and remodelling rather than being seen as a sequential process. (b) Effect of RVH-induced ischaemia. RVH is characterised by tissue hypoxia arising from ischaemia and microcirculatory insufficiency. Ischaemia-derived ROS, through the activation of transcription factors, drive the metabolic remodelling, contractile dysfunction and fibrosis that occur in RV failure. RVH, RV hypertrophy; PA, pulmonary artery; ROS, reactive oxygen species; MMPs, matrix metalloproteinases. (From Iacobazzi [21]. It is an open access article)

Also, in response to increased afterload, there is an activation of the foetal gene pattern in RV, re-expressing of genes from normal foetal RV. This includes a shift from α- to β-myosin heavy chain expression and an increase in adrenergic receptors, calcineurin activation [22,23,24], and phosphodiesterase type-5 (PDE5) expression [25]. The foetal gene pattern re-expression, particularly the myosin heavy chain shift from the α to β isoform, an hallmark of foetal gene reactivation, is also triggered in LV failure (LVF) [22].

Further, using microarray gene chip studies of mice, Urashima et al. compared LV hypertrophy (LVH) from aortic banding with RVH from pulmonary banding, and they demonstrated both similar and different LV and RV adaptive mechanisms [26]. One pathway that is more activated in the pressure-loaded RV compared with the pressure-loaded LV is the Wnt signalling pathway (Fig. 4.2) [27,28,29]. Wnt regulates glycogen synthesis kinase-3b activity, a serine/threonine protein kinase active in multiple intracellular signalling pathways, including cell proliferation, migration, inflammation, glucose regulation, and apoptosis [28, 29]. Also, there are multiple variation concerning the RV and LV in their adaptation to increased loading and likely differences in metabolism, mitochondrial remodelling, and glycolysis-to glucose oxidation coupling. These metabolic changes may subsequently lead to hyperpolarization of the mitochondrial membrane potential in RV hypertrophy, inefficient energy metabolism, and increased lactate production at an earlier stage of maladaptation compared with the LV [30].

Fig. 4.2
figure 2

The Wnt/β-catenin signaling pathway . In the Wnt-off state, defined by the absence of an active Wnt ligand, β-catenin is phosphorylated by the destruction complex (formed from the two kinases Gsk3 and Ck1, the scaffolding protein Axin, and the tumor suppressor Apc) and degraded by the ubiquitin-proteasome pathway. In the Wnt-on state, active Wnt ligands interact with the Fz receptors and the Lrp5/6 coreceptor. Phosphorylation of Lrp5/6 by Gsk3 and Ck1 recruits Dvl and Axin to the receptor complex and hence inhibits the destruction complex. This, in turn, inhibits β-catenin phosphorylation and stabilizes β-catenin in the cytoplasm. β-catenin is then translocated into the nucleus, by a complex including Fam53b/Smp, and regulates target gene expression with the Tcf/Lef transcription factors. Many modulators including the inhibitors sFrps and Wif are known to tightly regulate the signaling cascade. (From Ozhan et al. [27]. It is an open access article)

In CR process, several cell markers may indicate an undergoing CR progression, as well as alterations with an rise in α- and a reduction in β-myosin heavy chain, raised exhibition of Glucose transporter type (GLUT)-1, α-actin, natriuretic peptides, galectin, caveolin, neuronal nitric oxide synthase (NOS), angiotensin-converting enzyme (ACE), reduction of GLUT-4, sarcoplasmic/endoplasmic reticulum Ca(2+)ATPase 2a (SERCA 2a), and a change from free fatty acids oxidation to glucose metabolism [31, 32].

To sum up, cardiac dysfunction is the most important effect of CR. Because of cardiac injury, CR begin with genetic alterations, with reexpression of foetal genes, with cellular and molecular modifications, and gradually damage of ventricular function that develops with signs and symptoms of HF (Fig. 4.3) [4, 31, 33,34,35].

Fig. 4.3
figure 3

Sequence of events from cardiac injury to cardiac dysfunction. (From Azevedo [33]. It is an open access article)

2 Adaptive Versus Maladaptive Cardiac Remodelling

As already stated, CR is an adjusting and a maladaptive process. The adjusting process sustains heart function due to pressure overload or volume overload in case of acute cardiac injury [36]. Even if CAD affects directly RV with regional or global ischemia, RV physiology and RVF are mainly influenced by raised preload or afterload [20]. It has to be underlined that RV is exclusively dependent on afterload. Even small changes in total PVR, as demonstrated by modest increases in mean airway pressure during positive pressure ventilation, can reduce RV contractile performance and lower CO even when RV preload is maintained [37]. In contrast, significant changes in LV afterload may induces only modest changes in LV stroke volume [38]. Although patients with acute changes in systemic vascular resistance can compensate over a wide range, those with acute pulmonary arterial hypertension (PAH) if associate acute lung failure, often develop overt RVF and compromised CO [20]. In the largest part of clinical scenarios, even acute mild/moderate raises in RV afterload produce significant falls in RV output, including PAH and RV outflow obstruction, with the mention that usually changes in afterload are chronic and occur progressively [20]. Undeniably, in the chronic conditions, the relative increase in RV afterload is much greater in PAH than the increase in LV afterload in systemic hypertension [20].

Only that, long-term CR is damaging and correlated with a weak prognosis [39, 40]. Significant CR causes the increase of failing cardiac function [39, 41] that is related with bad prognosis especially for MI [40]. Actually, there is no evidence to suggest the time of occurrence from adaptive remodelling to maladaptive remodelling or if CR can be recognized at right moment. For instance, continuing CR is usual after an initial, moderately large anterior MI, but is uncommon after an initial small inferior MI [41]. CR after acute MI involving the LV may progress with LV dilatation and with later RV dilatation. Biventricular (BiV) remodelling comprises a group of patients with extremely poor outcomes [42]. Importantly, BiV failure is regarded as the terminal phase of CR [43, 44]. On the other hand, there is no clear knowledge regarding the effects of acute MI of LV and RV remodelling. It seems that the most important pathophysiological mechanism is the PH followed by raise of the RV afterload.

It should be mentioned that about 50% of patients with cardiac failing will die in five years. Moreover, about 40% of patients with HF die during first one year of hospitalization [45]. Also, an important number of deaths related with CR and cardiac failing are produced by sudden death [46] suggestive of the fact that an asymptomatic patient doesn’t mean a convinced good prognosis. In the face of raised survival with up-to-date current treatments, death rates have inadmissible values [47].

3 Basic Concepts of Cardiac Remodelling

As already mentioned that “CR may be characterized as genome expression, molecular, cellular and interstitial changes that are manifested clinically as changes in size, shape and function of the heart after cardiac injury” [4], the myocyte or cardiomyocyte (CM) is the most important cardiac cell implicated in the CR process. Conversely, the various explanations of CR share common molecular, biochemical, and mechanical pathways. Furthermore, the interstitium, fibroblasts, collagen, coronary vasculature, hemodynamic load, and neurohumoral activation affect the process of CR. A shortly review is made with updated information for all factors that influence CR process especially in case of right heart.

3.1 Functional Changes

RV and LV are different in their anatomy and physiology. Moreover, morphologically and functionally, both ventricles are comprehensible linked not only in health but also as they react to disease [20]. In same time, alterations in ventricular mass, and changes in composition and volume negatively modify the cardiac function [39, 41, 48,49,50]. As CR continues over time, the heart dilates and becomes spherical instead of elliptical form [51, 52], with thinning of cardiac walls and mitral valve incompetence. Even so, the evolution of CR depends on the primary disease, the severity of the underlying disease, genotype, intermittent ischemia episodes, neuroendocrine activation, and recommended treatment [41, 53, 54].

3.2 Cellular and Molecular Changes

It should be emphasized that CR is related with numerous cellular changes as well as myocite hypertrophy, deficit of myocytes due to apoptosis [55,56,57] or necrosis [58], fibroblast proliferation [59] and fibrosis [60, 61]. At molecular level, recent literature has highlighted differences between the RV and LV in the expression of genes involved in the response to pressure loading and failure [62]. Some of these differences are detailed in the following text and are summarized in Table 4.1 [20, 25, 27, 63,64,65,66,67,68,69, 71, 72].

Table 4.1 Molecular differences between the left and right ventricles in response to adverse loading

3.3 The Cardiomyocyte (CM)

Human myocardiums are composed of myocytes tied and hold up by connective tissue mainly created from fibrillar collagen. The adult human heart have about 4–5 billions CMs but the myocardium has insignificant basic regenerative capability, and the damage of an important mass of cardiac muscle causes scar. In fact, the normal myocardium consists of four components that are highly interrelated: CMs, CFs, the microcirculation and the extracellular matrix (ECM) [73]. All four above components have decisive role in the progression of chamber remodelling with hypertrophy [73]. RV myocytes have mainly longitudinal myocyte direction with angulated intrusion of superficial myocytes toward the endocardium creating a peristaltic contraction from the inlet to outlet and a bellows-like motion of the free wall toward the septum [74]. In addition, RV myocytes present quicker twitch velocities than LV myocytes [75].

Even if the RV and the LV have related cellular and molecular responses to stress, there are various distinctions at the cellular and molecular levels in their responses to stress such as pressure overload. Furthermore, both ventricles show comparable modifications in genes controlling ECM and cytoskeleton remodelling, but with significant differentiation in genes controlling energy production, mitochondrial function, reactive oxygen species (ROS) production, antioxidant protection, and angiogenesis [26, 70].

Unlike the CMs that comprise almost 1/3 of all heart’s cells [76], endothelial cells (ECs) [77], vascular smooth muscle cells (VSMCs) [77], CFs, macrophages and surrounding ECM that exist in the cardiac interstitium are together named as nonmyocyte cells [78]. The development of nonmyocyte cells is mentioned as interstitial structural remodelling and is characterized by the increase of collagen [78, 79]. Because the increase of nonmyocytes and myocytes is unrelated of each other, the hypertrophic process may be a similar and proportional or heterogeneous with excessive nonmyocytes raise, correspondingly [80, 81].

Since CMs have low ability for cellular multiplication, it is clear that they can growth by cellular enlargement. Consequently, the cross-sectional area and diameter of CMs are raised. Furthermore, typical characteristics of hypertrophy develop too: (1) sarcomere is intense restructured; (2) raised CMs size and myocardial mass by boost of protein synthesis; and (3) cardiac specific gene expression suffers alterations [11, 82]. A part of these modifications are known as re-activation of foetal gene program that implies only the re-expression of normal genes from embryonic and neonatal heart, together with contractile foetal proteins such as skeletal α-actin, atrial myosin light chain-1 and β-myosin heavy chain, and signal transduction proteins such as atrial natriuretic peptide (ANP) or B-type natriuretic peptide (BNP) [9, 83]. Abnormal existence of these foetal proteins in adult human heart has an effect on cardiac contraction, myocardial metabolism including Ca2+ control, resulting in maladaptive CR [84]. Of particular interest, the presence of the foetal gene program does not exist in physiological hypertrophy [85].

Changes of CMs in the size, shape, and function are related with the raise of cell death as well. Deficit of CMs is mainly related to the chronic CR process with progression to HF, increased apoptosis [86], and decreased cardiac function of heart. For that reason, the equilibrium between CMs survival and apoptotic pathways seems the mainly factor of the shift process from hypertrophy to ventricular dilatation [87].

It has previously described that the term matricellular proteins don’t have any significant role in cardiac tissue structure, but they are stimulated by injury with the alteration of cell to cell and cell to matrix connections [88]. Therefore, the production of matricellular proteins in the cardiac matrix, causes their attachment to growth factors, cytokines, and cardiac cells receptors of transducing signalling cascades. Consequently, matricellular proteins are controlled in CR process and have a significant function in controlling of inflammatory, reparation, fibrotic and angiogenic process [88]. It should be remembered that the term matricellular protein has been created by Bornstein [89] for ECM proteins which have no involvement in the structure of the ECM, except they appear and control cardiac cellular function subsequent to injury. According to current evidence, matricellular proteins implied in CR comprise secreted protein acidic and rich in cysteine (SPARC), osteopontin, thrombospondin (TSP), periostin, and tenascin families [88, 90]. Also, initiation of the matricellular proteins (TSP-1, tenascin-C, SPARC) can produce the process of “de-adhesion” in tissue remodelling. This process may be significant in supporting cell motility during inhibition of cell anoikis [88].

To sum up, the normal adult heart contains CMs, a complicated system of ECM, and nonmyocytes that are more numerous than CMs. Every CM is encircled by collagen (endomysium), and connective tissue (perimysium) demarcates individual fibbers. Also, the normal mammalian heart has a rich vascular system consisted of capillary, venous and arteriolar endothelial cells, pericytes, smooth muscle cells, numerous CFs, minor numbers of macrophages, mast cells, lymphocytes and dendritic cells [88]. After an injury of sufficient degree, CMs diminish and develop into elongated or hypertrophied cells to sustain stroke volume [49, 50]. As well as, the width of ventricles wall may raise due to myocyte hypertrophy [48,49,50]. Changing of heart loading conditions like raised preload further causes stretching of cell membranes and increases of wall stress both have the ability to initiate the effect of hypertrophy genes. More precisely, in cardiac myocytes may be triggered new contractile proteins synthesis joining with new sarcomere. The final result is believed to be CMs elongation or the increase of their diameter [91].

It seems that RV dysfunction is described in PAH or RV obstruction [21]. Patients with RV dysfunction are put together even though the CHD can initiate different molecular, cellular and functional remodelling in the RV [21]. In addition, the evaluation of RV function is mainly based on techniques which assess structure and function of RV (e.g. echocardiography, MRI and pulmonary angiography) instead of exploring the cellular and the molecular irregularities of the RV dysfunction from CHD [21]. For instance, some studies demonstrated the alteration of gene expression in signalling pathways that control heart growth in children with Tetralogy of Fallot (TOF). Modifications such as significant suppression of genes in the Notch and Wnt pathways, in VEGF gene expression and numerous ECM proteins are identified as factors that lead to TOF [92, 93]. Another genome-wide array study has demonstrated obvious difference in gene expression between the TOF and other RVH phenotypes, including VSD and ASD. Genes related with cardiac maldevelopment such as SNIP, A2BP1 and KIAA1437 are more active in the TOF group, and genes linked with stress reaction and cell proliferation are more exhibited in the RVH conditions [94].

In addition, a molecular conversion from RV to LV characteristics appears for the period of RV adjustment to pressure overload, with the mention that altered genes from RVH have a normal representation same to the normal LV tissue. Additionally, the association of tissue hypoxia and hypertrophy can boost the protein phosphatase PP1 activity leading to raised phospholamban (PLN-Ser16) dephosphorylation in CMs, followed by cardiac dysfunction [95]. As already described, hypoxia-inducible factor-1 (HIF1α) is a further contributor factor in the RV adjustment to tissue hypoxia and mechanical stress. In acute hypoxia, HIF1α is cardioprotective based on its property to produce angiogenic, metabolic and erythropoietic genes [96]. Conversely, HIF1α sustain transforming growth factor beta TGFβ1-mediated organ fibrosis in chronic hypoxic states [96]. Also, genetic differences of HIF1α change myocardial adjustment to hypoxia during postsurgical period and before RV remodelling process [97]. Therefore, the adjustment of RV to hypoxia prior to TOF surgery is based on the HIF1α pathway and could have an effect on RV phenotype after surgery.

3.4 Matricellular Proteins: TSP Family

It should be underlined that one of the matricellular proteins, TSP family contains five members divided into two groups. TSP-1 and TSP-2 form homotrimers (Subgroup A), while TSP-3, TSP-4 and TSP-5 (COMP) form homopentamers (Subgroup B) [73, 98,99,100] (Figs. 4.4 and 4.5).

Fig. 4.4
figure 4

Structure of the two thrombospondin (TSP) subgroups . Subgroup A form homotrimers and consist of TSP-1 and TSP-2, while Subgroup B form homopentamers and consist of TSP-3, TSP-4, and TSP-5 (COMP). Subgroup A has domains that bind to CD36 and inhibit MMPs. The N-terminal domains tend to be family member specific, while the CTD has high homology between the family members. NTD N-terminal domain (specific to each family member), vWF-C vonWillebrand factor C-type domain, MMP matrix metalloproteinase, EGF epidermal growth factor, CTD C-terminal domain. (From Kirk et al. [73] with permission)

Fig. 4.5
figure 5

Recent work has identified several roles for TSPs once the heart has transitioned to heart failure. The effects of the TSPs on matrix remodeling and inflammation are still important in HF. However, they also regulate a number of pathophysiologically important elements within the cardiac myocyte, which exhibits hypertrophy, apoptosis, and contractile dysfunction with HF. However, our knowledge of the mechanistic function of each of the TSPs in the HF myocyte is still incomplete. (From Kirk et al. [73] with permission)

These matricellular proteins induced by heart injury with CR named TSPs have a significant function during cardiac growth [101]. However, numerous TSP proteins activity increases to stress. They have the capacity to attach with the components of the ECM such as cytokines and growth factors. In addition, it’s convenient at this point to discuss that pressure overload quickly raise TSP expression, mainly TSP-1 and TSP-4 [102, 103], and volume overload shows a significant raise in TSP-4 mRNA [104]. In most of the cases, the mechanisms implied are same to the post-MI, as well as inflammation [105] and fibrosis [106].

3.5 Endoplasmic Reticulum Stress

Theoretically, stress factors such as hypoxia, ischemia/reperfusion, hypertrophy, pressure overload, and drug-induced insults can cause activation of endoplasmic reticulum (ER) stress in the heart [107]. ER stress being closely implied in the protection of cardiovascular homeostasis, it proves a significant therapeutic aim for cardiovascular diseases treatment. Figure 4.6 shows a schematic illustration of ER stress pathways with particular highlighting on their role in cardiac physiology and pathology [108]. The ER stress response or ‘unfolded protein response’ (UPR) is essential for normal cellular protection, but in CR like HF can generate apoptosis [107, 108]. A simplistic explanation is TSP-1 and TSP-2 have anti-inflammatory, pro-fibrotic, and antiangiogenic properties, as TSP-4 induces pro-inflammatory and pro-angiogenic consequences [109]. Also, the TSPs are significant aims for stopping the evolution from MI to HF.

Fig. 4.6
figure 6

ER stress signaling pathways. (a) PERK-dependent pathway activated by ER stress. PERK, a transmembrane kinase and endoribonuclease, interacts with BiP/GRP78 under nonstressed conditions. On activation of ER stress, BiP/GRP78 dissociates from PERK, resulting in dimerization of PERK and activation of its kinase domain, autophosphorylation, and subsequent phosphorylation of eIF2α. Phosphorylation of eIF2α results in attenuation of protein synthesis. However, expression of ATF4 is not inhibited, and the transcription factor induces expression of ERSR-containing genes. (b) ATF6 pathway. Under nonstress conditions, ATF6, a transmembrane protein localized to the ER, interacts with BiP/GRP78 and calreticulin. After ER stress, BiP/GRP78 and calreticulin dissociate from ATF6, and the protein translocates to the Golgi, where it undergoes cleavage by S1P and S2P proteases. This cleavage yields a cytoplasmic transcription factor (N-ATF6) that translocates to the nucleus and induces ERSR-containing genes. (c) IRE1 pathway. IRE1 is an ER transmembrane protein containing a serine–threonine kinase domain and a carboxyl-terminal endoribonuclease domain in its cytoplasmic region and binds to BiP/GRP78. Under ER stress conditions, BiP/GRP78 is released from IRE1 followed by IRE1 homodimerization and autophosphorylation. Phosphorylation is essential for IRE1 endoribonuclease activity that is responsible for splicing of XBP1 mRNA, yielding spliced XBP1s mRNA encoding a potent transcription factor. The XBP1s splice variant binds to ERSE-containing promoters and activates ERSE genes. XBP1s also binds to a second cis-acting motif, termed the UPRE, resulting in upregulation of genes involved in ERAD. (From Groenendyk et al. [108] with permission)

3.6 Pleiotropic Functions of Cardiac Fibroblasts (CFs)

In particular, CFs include over 50% of the cells in the adult heart [110] being implied in the development and deterioration of the cardiac ECM by generating collagens, proteoglycans, MMPs and TIMPs. As well, CFs produce different bioactive mediators such as VEGF-A, fibroblast growth factors (FGFs), transforming growth factor beta (TGF-β), platelet-derived growth factor (PDGF) which affect cardiac angiogenesis and CM proliferation. Furthermore, CFs have an effect on cardiac electrophysiology by protecting CM bundles, spreading of electrical signals, and changing mechanical stimuli in electronic signals [111]. In fact, CFs develop intracellular electrical coupling and interconnect with CMs through gap junctions (Fig. 4.7) [112,113,114]. Also, CFs being the most abundant cardiac cell type, they monitor CM proliferation during heart development. Therefore, CFs activation is a critical early repair response after cardiac injury.

Fig. 4.7
figure 7

Cell Communication Between Cardiac Fibroblasts and Myocytes . Z-section of a cell aggregate containing cardiac fibroblasts that were dual loaded with Lucifer Yellow and CMRA and myocytes that were unloaded. Cardiac fibroblasts appear as yellow or orange cells. The orange cells are fibroblasts that have transferred their green dye to an adjacent cell (orange arrow). Note the green cells, which are myocytes that have received green dye from an adjacent fibroblast (white arrow). (From Souders et al. [112] with permission)

3.7 Collagen Synthesis and Degradation

Collagen is synthesized by interstitial CFs which are degraded by locally formed enzymes named collagenases, such as matrix metalloproteinases (MMPs). The cardiac interstitium consist of 95% of type I and type III collagen fibers. The most important roles of collagen network are to control apoptosis, fix pathological processes, preserve the configuration of structures, control the resistance conduction during fiber shortening, and produce cytokines and growth factors [115]. Each heart has inactive myocardial collagenases in the ventricles but they are activated after a myocardial injury [116]. As already noted above, in any failing heart, CR primarily arises as a preventing reaction to protect the myocardium structure, but with gradually collagen deposit can causes cardiac fibrosis with diastolic and systolic dysfunction [117,118,119,120]. Importantly, collagen XIV is necessary to produce and preserve the ECM network in the heart growth [121]. Throughout fibrosis process, CMs undertake hypertrophic modifications, while MyoFb continue with collagen production and scar formation at the site of injury. In addition, collagen XI is necessary for myocardial growth supporting the nucleation of type I and II fibrils [122]. Therefore, the increased activity of collagen type XI alpha 2 chain (COL11A2) gene can be correlated with production of heterotypic fibrils with collagen I that is implicated in CR [123]. Nevertheless, persistence of CFs in injury area leads to chronic scar and remodelling [119].

According to evidence, the atypical deposit of type III collagen and type I collagen was discovered in cardiac injury, produced by several signalling pathways such as TGF-β, endothelin-1 (ET-1), angiotensin II (Ang II), connective tissue growth factor, and PDGF. In this situation, fibrosis is related with raised myocardial stiffness, decreased diastolic function, reduced contraction, failed coronary flow and malignant arrhythmias [124, 125].

As a result, collagen has an important role in the protection of cardiac structure and function. In CR, the equilibrium between collagen production and degeneration is altered with numerous side effects. To prevent developing of CR to HF is necessary of timely developed stable scar that restores the injured tissue [126]. This equilibrium is preserved to some extent by MMPs that alter the ECM and tissue inhibitors of MMP (TIMPs) [127, 128]. In case of post MI, numerous mechanisms are run by CFs that change into MyoFb [129].

It should be noted, that there is a direct support between TSP-1 and the cardiac collagen. More important, TSP-1 and TSP-2 can preserve ECM normal structure by controlling the MMPs [73, 130, 131] (Table 4.2). TSP-1 and TSP-2 can directly attach to MMP2 and MMP9 through their Type 1 Repeats [132], but this attachment doesn’t cause their inhibition directly [133]. When TSPs perform their activation by TGFβ, this diminishes MMP transcription [134]. Also, TSP-1 can attach to collagen V and fibrinogen [135, 136] and hasten fiber growth [137]. On the other hand, TSP-4 increases fibrosis by production of collagens I, II, III, and V [138] that support its direct role on the ECM remodelling process [139]. Essentially, in order to regulate cardiac fibrosis, TSP-4 expression is controlled by the transcription factor Krüppel-like factor 6 (KLF6) [140]. Also, TSP-4 triggers TGF-β [73] necessary for the transdifferentiation of MyoFb from CFs [141]. Further, MyoFb are indispensable to the cardiac fibrosis by production of collagen [142].

Table 4.2 Thrombospondin binding partners . From Kirk et al. [73] with permission

3.8 Apoptosis

There are three main mechanisms involved in myocyte death: apoptosis or programmed cell death, necrosis and autophagy. According to data, cardiac dysfunction is correlated with modifications produced by autophagy, that can be adaptive or deleterious [143,144,145]. Initially, Sharov et al. have been suggested that raised cardiac apoptosis with CMs damage increases LV dysfunction with chronic HF [55]. Conversely, Olivetti et al. showed on myocardial samples from patients who underwent heart transplantation that cardiac apoptosis was increased more than 200-fold in the patients with failing heart [57]. In general, apoptosis has an important function in cardiac growth and in various heart diseases with ischemic and non-ischemic origin [146, 147]. However, the major mechanism of CM death from MI is the coagulation necrosis, even if apoptosis is also implied in CMs damage (Fig. 4.8) [148, 149].

Fig. 4.8
figure 8

Typical appearance of different types of cell death. (a, c, e) Confocal micrographs: counterstaining for actin, red; nuclei, blue; specific fluorescence, green. (b, d, f) Electron microscopic pictures (all bars = 2 μm). (a, b) Apoptotic cell death . (a) Nuclei with DNA fragmentation are green. (b) Nuclei show condensed chromatin. (c, d) Oncotic cell death. (c) Single cell oncosis labeled with C9. (d) Nuclei are electron-lucent with clumped chromatin, mitochondria are damaged with flocculent densities. (e, f) Autophagic cell death. (e) Ubiquitin deposition and loss of nuclei. (f) Ultrastructural appearance with numerous autophagic vacuoles. (From Kostin et al. [148] with permission)

It is important to note, that there is a fast triggering of caspase-3 in MI during 1 h after the onset of ischemia [150, 151], and the process of CM apoptosis can be completed within 24 h (Figs. 4.9 and 4.10) [42, 152, 153]. Previous reports have been shown that apoptosis is associated with unfavourable CR and HF post-MI in case of ablated proapoptotic protein Bnip3 [154, 155]. Also, TSP-1 and TSP-2 trigger apoptosis in ECs from microcirculation [156, 157]. Apoptosis inhibits endothelial tubule development and consequently has antiangiogenic effect. Further, vitro studies showed that matricellular proteins relate with the transmembrane glycoprotein CD36 [157,158,159,160,161,162], being the mainly mechanism that causes the antiangiogenic effect. Also, Primo et al. showed in cultured human ECs the inhibition of angiogenesis by the pathway of the VEGF receptor modulated by TSP-1 [163]. Attachment of TSP-1 to CD36 causes apoptosis of ECs by raise of death receptors and Fas ligand [164].

Fig. 4.9
figure 9

Schematic depiction of pathways leading to programmed cardiomyocyte death , as described in the text. Mechanisms of cell death (bottom) are, from left to right, caspase-dependent apoptosis, caspase-independent apoptosis, programmed necrosis, and authophagy. Solid lines show primary effects; interrupted lines depict cross-talk between pathways. (From Dorn [152] with permission).

Fig. 4.10
figure 10

Demonstration of apoptosis . (a) Labelling of nuclear DNA fragmentation using terminal deoxynucleotidyl transferase mediated dUTP nick end labelling (TUNEL). Colocalisation of (b) TUNEL and actin and of (c) TUNEL and activated caspase 3. (From Bussani et al. [42] with permission)

ADAMTS-7 is a member of the disintegrin and MMPs with TSP motifs (ADAMTS) family [73, 165], being newly recognized to be considerably correlated genome-wide with angiographic CAD [73, 166]. It has been previously confirmed that ADAMTS-7 support VSMCs migration and post injury neointima production via degradation of the matrix protein ‘Cartilage Oligomeric Matrix Protein’ (COMP or TSP-5) [73, 167]. Same researchers demonstrated that ADAMTS-7 breaks down the TSP-1 with re-endothelialization prevention [73, 168]. Also, the extracellular proteases thrombin, cathepsins, leukocyte elastases and plasmin can degenerate the TSPs [169].

Autophagy is defined as the intracellular process that removes needless cytoplasmatic components by lysosomes [144]. To date, the ubiquitin-proteasome system (UPS) and the autophagic-lysosomal pathway (ALP) are two major pathways in charge for most cellular proteins deterioration. Modifications of UPS and ALP pathways are correlated with the increase of proteotoxic defective proteins in the heart, a characteristic of frequent heart disease [144]. Acute ALP inhibition (proteasome inhibition) boost occasionally ‘intrinsic proteasome peptidase activities’, but chronic ALP inhibition blocks UPS pathway functioning in ubiquitinated protein stage [144]. As a result, autophagy has a significant function in proteotoxicity prevention by the ubiquitin system [144], and chaperones (heat shock protein-HSP) [145]. Particularly, the co-chaperones Bag3 and HspB8 have significant role in the heart autophagy by chaperone-assisted selective autophagy [170, 171]. Regardless of myocyte death, the gradual decrease of CMs thas a significant function in CR and could be a potential target for therapeutic interventions.

4 Fibrosis

Just as RV fibrosis is commonly seen in the setting of both severe RV afterload and chronic pulmonary regurgitation, LV fibrosis is common in both aortic stenosis and regurgitation [172,173,174]. At the site of MI, acute focal fibrotic scarring provides myocardial healing and prevents rupture [175]. In contrast, chronic diffuse or focal reactive myocardial fibrosis is a result of either pressure overload or volume overload due to persisting hypertension, metabolic disorders, valvular heart diseases, ischemic injury (in areas remote from the infarction), or diffuse myocardial diseases, such as cardiomyopathies [175].

Myocardial fibrosis is defined by dysregulated collagen turnover characterized by increased synthesis that dominates over unaffected or reduced degradation [176, 177] with excessive diffuse collagen accumulation in the interstitial and perivascular spaces [178]. For that reason, the dysregulation of distinct pro- and antifibrotic factors, including cytokines and chemokines, growth factors, proteases, hormones, and ROS, is responsible for the alteration of the collagen matrix (Fig. 4.11) [179, 180].

Fig. 4.11
figure 11

Schematic representation of biochemical and cellular mechanisms of cardiac fibrosis . Under physiological conditions (left), fibroblasts secrete extracellular procollagen chains into the interstitium that assemble into fibrils and are cross-linked by lysyl oxidase. Several cell types are implicated in fibrotic remodelling of the heart either directly by producing matrix proteins (fibroblasts), or indirectly by secreting fibrogenic mediators (macrophages, mast cells, lymphocytes, cardiomyocytes, and vascular cells). Under pathological conditions (right), alterations in the matrix environment, induction and release of growth factors and cytokines, and increase of mechanical stress dynamically modulate fibroblast transdifferentiation into myofibroblasts. Higher collagen cross-linking results in increased myocardial tensile strength. Resistance to degradation by matrix metalloproteinases (MMPs) increases cross-linked collagen, which favours matrisome expansion. Pink, grey, and green boxes list part of the secretome of mycocytes, myofibroblasts, and macrophages/leucocytes/mast cells, respectively, that trigger and maintain fibrosis. Gal-3 galectin-3, IL interleukin, PDGF platelet-derived growth factor, RAAS renin–angiotensin–aldosterone system, ROS reactive oxygen species, TGF transforming growth factor, TNF tumour necrosis factor. (From Gyöngyösi et al. [179]. It is an open access article)

The degeneration of collagen turnover takes place mainly in phenotypically transformed fibroblasts, termed MyoFb [79, 181]. The shift of CFs in MyoFb implies the expression of α-SMA, a characteristic of SMCs [79, 181,182,183,184,185,186]. As well as, the development of a wide active ER stimulated by a number of bioactive effectors [79, 181,182,183,184,185,186]. CFs and particularly MyoFb form collagen type I and III fibrils and develop into cross-linked to form the final fibres [176]. Collagen cross-linking is a significant post-translational stage that raises the resistance of collagen fibres to degradation by MMPs [187, 188]. Only that, myocardial fibrosis disrupts the myocardial architecture, contributes to myocardial disarray, and determines mechanical [189], electrical [190, 191] and vasomotor [192] dysfunction, thus promoting the progression of cardiac diseases to HF [175]. Fibrosis is induced by various genetic disorders, pressure or volume stress, heart injuries, and other diseases. There is evidence that depending on the particular trigger, distinct molecular pathways have varying importance for the individual types of fibrosis. As the development of myocardial fibrosis is characterized by a complex dysregulation of a number of different factors including inflammatory chemokines, angiotensin II (Ang II), and endothelin signalling, the FIBROTARGETS consortium that is a multinational consortium with industrial and academic partners, funded by the European Commission is primarily aimed for characterizing novel emerging mechanisms of myocardial fibrosis [180]. Targets and biomarkers under investigation include especially proteins, proteoglycans, and microRNAs (miRNAs) [180].

In increased volume loading, the RV appears more prone than the LV to develop fibrosis [63]. Similarly, patients after surgical repair of TOF who have long-standing RV volume load secondary to pulmonary insufficiency develop RV fibrosis [172]. This is clinically important as risk factor for increased propensity to arrhythmias, exercise intolerance, and RVF [172, 193]. It has been suggested that these differences in response between the RV and LV to volume loading may stem from the different embryological origin of the two ventricles [63].

Several single or multimodal imaging technologies have been used to assess the extent and type of myocardial fibrosis. Besides the direct morphological display of the fibrotic tissue, indirect cardiac functional imaging may evidence fibrosis correlated with decrease of systolic function and increased myocardial stiffness with diastolic dysfunction [175]. Cardiac magnetic resonance imaging (MRI) provides detailed tissue characterization, identifying focal myocardial fibrotic scars with late gadolinium enhancement (ventricular LGE) and an estimation of diffuse myocardial fibrosis with post-contrast enhanced T1 and T2 mapping (Fig. 4.12) [179, 194].

Fig. 4.12
figure 12

Representative native and T1 cardiac magnetic resonance imaging (cMRI) of diffuse myocardial fibrosis. (a) Diffuse myocardial fibrosis on the short-axis view of the cMRI image, with the circumference of the anteroseptal myocardial area (region of interest). (b) cMRI T1 map of a patient with moderate aortic stenosis and moderate diffuse myocardial fibrosis. (c) cMRI T1 map of another patient with severe aortic stenosis and severe diffuse fibrosis of the left ventricle. (From Gyöngyösi et al. [179]. It is an open access article)

Positron emission tomography (PET) imaging performed by using 15O-labelled water (H2 15O) and carbon monoxide (C15O) allowed the non-invasive quantification of both myocardial perfusion and fibrosis [195]. Combining PET and MRI has the potential for sensitive and quantitative imaging of cardiovascular anatomy and function with detection of molecular events at the same time [196, 197]. It’s worthwhile to specify that PET–MRI (Biograph mMRI, Siemens AG) image allows the simultaneous detection of myocardial global and regional function, ECM volume, and tissue perfusion and metabolism [198].

Histopathological analysis of endomyocardial biopsy specimens is the current gold standard for diagnosis and assessment of cardiac fibrosis. A number of circulating biomarkers, including (pro-) collagen cleavage products, processing enzymes, but also miRNAs (Table 4.3), have been proposed and analysed [179]. Details of these biomarkers and potential targets have been described previously including proteins and proteoglycans that impact fibrosis and miRNAs that act in fibrosis [180]. For their use as cardiac fibrosis biomarkers, it seems reasonable that a combination of several from these increases the predictive power, particularly in the case of miRNAs [199, 200].

Table 4.3 Potential circulating biomarkers for assessment of cardiac fibrosis

As a consequence, the treatment of HF patients improves clinical symptoms, but does not reverse fibrosis. Furthermore, the severity of histological proven myocardial fibrosis has been reported to be associated with higher long-term mortality in patients with cardiac diseases, mainly patients with HF [200, 201].

4.1 miRNAs

Genetic variations exist among the RV and the LV. Drake et al. [71] note the dissimilarity between gene expression patterns in normal RV and LV in both mRNA and microRNAs (miRNA) types. More precisely, the transcription factor Irx2 is not exhibited in the RV but insulin-like growth factor 1 (IGF-1) is exhibited mainly in the LV. Moreover, same team made the assumption that these dissimilarities can be the result of different embryologic origin or the RV is a low-pressure chamber compared to the LV. Also, Reddy et al. [66] demonstrated firstly that changes in miRNAs exist in RV remodeling from RVH to RVF and are mostly comparable to pressure-stressed LV but with separate signalling regulatory pathways.

RV dysfunction is described entirely in RV obstruction or PAH [21]. Only that, all patients with RV dysfunction are put together regardless of the fact that CHD have various functional, molecular and cellular remodelling patterns in the RV [21] (Fig. 4.13). Blood biomarkers, similar to plasma proteins and miRNAs represent an important way to evaluate the function and remodelling of RV [21]. Heart miRNAs are constant and quantifiable discharged in the blood flow as exosomes, microvesicles or joining with high-density lipoproteins (HDL) and RNA-binding proteins [21].

Fig. 4.13
figure 13

Dysregulated miRNAs in congenital heart diseases (CHDs) . A figure showing the link between CHD and miRNAs in cardiomyocytes. Small number of miRNAs are upregulated in cardiomyocyte during CHD. These miRNAs can be released from the cell in microvesicles, by incorporation into exosomes, by linkage to high-density lipoproteins or bound to RNA-binding proteins. Dysregulated levels of miRNAs, crucial in RV development, are found in the bloodstream of children with VSD. The differentially expressed has-miR-222-3p, has-let-7e-5p and has-miR-433 bind with specific transcription factors (NOTCH1, GATA4, HAND1 and ZFPM) associated with RV morphogenesis. (From Iacobazzi et al. [21]. It is an open access article)

There are numerous disrupted miRNAs during CR and RHF [21, 202,203,204,205] (Fig. 4.14). miRNAs are non-coding single-stranded RNAs formed from 19–24 nucleotides that adjust in the negative way the exhibition of a particular mRNA via translational degeneration or suppression [206]. According to data, there have been shown in children with VSD in comparison with controls eight various miRNAs. Particularly, NOTCH1 is implied in ventricular growth, and GATA4 has an important function in atrial and ventricular growth, heart partition, and atrioventricular valve development [21, 207].

Fig. 4.14
figure 14

Functional role of miRNAs in the normal and diseased heart. A normal and a hypertrophic heart are shown in schematic form, depicting miRNAs that contribute to normal function or pathological remodelling. The expression of selected miRNAs within the heart is shown, along with their corresponding functions. All arrows denote the normal action of each component or process. miR-1 and miR-133 are involved in the development of a normal heart (left) by regulating proliferation, differentiation and cardiac conduction. For example, proliferation is promoted by cell-cycle regulators, but miR-1 and miR-133 block these regulators, thus blocking proliferation. miR-208a also contributes to the regulation of the conduction system. After cardiac injury (right), various miRNAs contribute to pathological remodelling and the progression to heart failure. miR-29 and miR-21 block and promote cardiac fibrosis, respectively. miR-29 blocks fibrosis by inhibiting the expression of ECM components, whereas miR-21 promotes fibrosis by stimulating mitogen-activated protein kinase (MAPK) signalling. miR-208 controls myosin isoform switching, cardiac hypertrophy and fibrosis. miR-23a promotes cardiac hypertrophy by inhibiting ubiquitin proteolysis, which itself inhibits hypertrophy. Hypoxia results in the repression of miR-320 and miR-199, which promote and block apoptosis, respectively. ECM extracellular matrix, LV left ventricle, MHC myosin heavy chain, RV right ventricle. (From Small et al. [205] with permission)

A low number of studies have studied the blood miRNAs in adult patients diagnosed with systemic RV [21]. Patients with the RV as the SV after transposition of the great arteries (TGA) had altered miRNAs profile. On the whole, from the 24 miRNAs various regulated, miRNA18a and miRNA486-5p related negatively with systemic ventricular contractility [21, 208]. Also, miRNA423_5p defined as a biomarker of LVF, has same expression in healthy adults and in SV after atrial repair of TGA adults [21, 209].

It seems that gene expression in signalling is changed in heart growth of children with TOF [21]. Alteration of VEGF gene expression and of a number of ECM proteins is established as contributors of TOF [92]. Important inhibition of genes in the Notch and Wnt pathways implied in heart growth are also found in children with TOF [21, 93]. Even if RVH is a component of TOF, there is a clear molecular difference between TOF and RVH gene expression, including VSD and ASD [21]. Whereas TOF children have unregulated genes for heart growth such as SNIP, A2BP1 and KIAA1437, RVH has a higher expression of genes implied in stress reaction and cell proliferation [21, 94]. What’s more, there is a molecular conversion from RV to LV features that appears during RV adjustment to pressure overload, as a result dysregulated gene phenotype from RVH is same with normal LV [21]. Another study based on cardiac tissues from RV in CHD [95], found dissimilar miRNAs in RV outflow tract obstruction (RVOT) in comparison with RVOT of normal infants [210]. Specifically, miRNA-424 and miRNA-222 had higher expression and they correlated with the decrease of heart growth being correlated with NF1 and HAS2 genes. Correspondingly, the increased expression of miR-421 in RV tissue from children with TOF is correlated with SOX4 gene necessary for cardiac outflow tract formation (Fig. 4.14) [21, 211, 212].

miRNA 133a is thought to suppress cardiac fibrosis and is decreased in LVF secondary to aortic constriction [213, 214]. This aligns with the marked upregulation of connective tissue growth factor/CCN2 and other profibrotic signalling molecules in the course of RV and LV fibrosis in models of RV afterload and RVF [71, 88, 215]. In contrast, miRNA 21 and 34c* may increase during LVF but decrease in RVF [71]. Reddy et al. [66] investigated miRNAs during the transition from RVH to RVF and compared these with miRNA expression in LVH or LVF. During RVH, there was altered expression of miRNAs 199a-3p, which is associated with CM survival and growth. With the progression to RVF and switching on the foetal gene phenotype, there was increased miRNA 208b, miRNA 34, miRNA 21, and miRNA1, which are associated with apoptosis and fibrosis [83]. These patterns of miRNA expression are largely related to LVH and LVF. Conversely, there are important distinction relating RV and LV miRNAs linked to cell survival, proliferation, metabolism, ECM production, and proteasome malfunction (miRNA 28, miRNA 148a, and miRNA 93), which were unregulated in RVH or RVF and downregulated or unchanged in LVH or LVF [66].

Common findings in both RVH and LVH are collagen deposition, fibrosis, and ECM remodelling [216]. The mechanisms inducing fibrosis are multiple, and in the setting of increased ventricular afterload, recognized triggers may include regional ischemia, necrosis, and apoptosis, among others [176]. There is an important match of the miRNA expression phenotype in human HF and foetal hearts in comparison with the adult normal heart tissue [204]. More studies data are necessary for a higher knowledge of these subcellular events that can guide to the development of new ventricle-specific treatments [217].

5 Other Factors

Factors that can also contribute to CR comprise endothelin, cytokines (tumor necrosis factor-alpha-TNFα and interleukins) [218], oxidative stress, MMPs, and peripheral monocytosis [219].

Endothelins (ET) are powerful vasoconstrictor peptides which increase in HF. The endothelin family of peptides is typically recognized for its vasoconstrictive properties. There are two known receptors for ET-1 in the heart, the ETA and ETB receptors, which have been shown to play differing and sometimes opposing roles. Importantly, ET-1 activation of the ETA receptor is known to increase collagen production in isolated human CFs [220]. Furthermore, MyoFb isolated from scar tissue after experimental MI have elevated levels of ET-1, suggesting an important function for ET within these cells [221].

ET-1 is a 21-amino acid peptide formed and discharged by the ECs and it has a quickest vasoconstrictive effect [222]. Cardiac ET-1 is active in both autocrine and paracrine effects by attaching to ETB receptors from cardiac ECs and ETA receptors from CMs [222]. The attachment of ET-1 to ETB receptors causes the discharge of signalling molecules such as NO and prostaglandin I2 [222]. If ET-1 attaches to the ETA receptors from CMs, it triggers CM constriction [222, 223]. As a consequence, there may exist a feedback mechanism concerning cardiac ECs and CMs that run CM constriction by the ET-1 system [222]. Also, patients with HF have raised exhibition of cardiac ET receptors and raised plasma ET-1 levels, both linked with disease severity [224]. ET antagonists they will be additionally efficacious in the treatment of pathological fibrosis in the heart [225]. Preliminary trials in humans had demonstrated beneficial hemodynamic and cardiac effects in patients with end-stage HF [226].

Cytokines (tumor necrosis factor-alpha TNFα and interleukins) are small peptides or glycoproteins that are discharged by nucleated cells [227]. Their temporary discharge adjust immune or repair processes by controlling cells growth, process of differentiating, metabolism, and protein synthesis [228]. Fibrinogen is an acute inflammatory regulator discharged by hepatocytes triggered by different cytokines. Also, CRP is an acute-phase reactant synthesized and discharged largely by hepatocytes in response to the cytokine IL-6. The highest levels of CRP are correlated with MI size but are reduced by early reperfusion [229]. It seems that IL-3 is a new biomarker of inflammation and can induce the multiplication of lymphocytes, macrophages, neutrophils, and monocytes with infiltration of heart where trigger the discharge of cytokines from CMs. Moreover, IL-3 can have significant functions in tissue repair. Understanding better inflammatory response could offer measurable ways of immune injury to tissues.

Leukocytosis was studied especially in MI [230, 231]. The ischemic-reperfusion stage produces the discharge of oxygen free radicals, cytokines, and other inflammation markers [231]. The presence of leukocytes in the microcirculation is followed by inflammatory reaction [232]. The transfer of leukocytes from blood flow to the vessel wall with tissue injury and inflammation is regulated by the selectin family of adhesion molecules with attachment of leukocytes to the ECs by involvement of integrins and diapedesis [233, 234]. Recruitment of leukocytes is mediated by complement triggering, TGF-β, IL-8, monocyte chemotactic protein-1 (MCP-1), and platelet activating factor (PAF) [235]. Also, the collection of neutrophils in the ischemic-reperfusion tissue could discharge proteolytic enzymes or ROS with further injury of myocytes. ROS directly injure CMs and vascular cells, and by triggering cytokines causes inflammation [236, 237]. Marginated neutrophils exert powerful cytotoxic effects through the adhesion with intercellular adhesion molecule-1 (ICAM-1) expressing CMs [235]. CD11b/ICAM-1 adherence activates the neutrophils respiratory burst resulting in myocyte oxidative injury [235].

Oxidative Stress produces important alteration of sarcolemmal and sarcoplasmic reticulum (SR) membrane, causing raise of intracellular Ca2+ levels with severe contraction of CMs, followed by mitochondrial damage and cell death [238, 239]. Specifically, ROS and redox signaling have an important function in apoptosis, including upstream signaling pro-apoptotic pathways and the mitochondria [240, 241]. There are signaling pro-apoptotic pathways that comprise the activation of ASK-1, JNK, p38MAPK, and CaMKII, as well as signaling anti-apoptotic pathways, such as Akt, Bcl2, and HSPs [241].

The cell resources of ROS comprise mitochondrial respiratory chain enzymes, xanthine oxidases (XOs), lipoxygenases, myeloperoxidases, uncoupled nitric oxide synthases (NOSs), and Nox proteins [242, 243]. Moreover, the important sources of ROS in the cardiovascular system comprise mitochondria, NADPH oxidases, NOSs, xanthine oxidases, cytochrome P450-based enzymes, and infiltrating inflammatory cells [243]. ROS are represented by free radicals (species with one or more unpaired electrons) such as superoxide (O2•−) and hydroxyl radicals (OH•), and nonradical species such as hydrogen peroxide (H2O2) [243]. In healthy adults, production of ROS is inhibited by enzymatic and nonenzymatic antioxidant systems that decrease ROS levels with preserving of a right redox balance in cells and tissues [243].

The first report of the presence of NADPH oxidases in human myocardium is of Heymes et al. [244]. NADPH or NADH-dependent ROS-generating activity are existent in nonphagocytic cell types [243], including VSMC [245, 246], ECs [247, 248], adventitial and CFs [249], and CMs [250]. Noxs are multi-subunit transmembrane enzymes that use NADPH as an electron donor to decrease oxygen to superoxide anion (O2−) and hydrogen peroxide (H2O2) [243]. Firstly, Noxs were described in phagocytes with the description of the Nox2 isoform that it also named gp91phox) [243, 251]. The rest comprise 6 other family members each coded by dissimilar genes, identified as Nox1, Nox3, Nox4, Nox5, dual oxidase 1 (Duox1), and Duox2 [243, 252,253,254]. All forms of Nox proteins demonstrate 21–59% similarity to Nox2, from which Nox3 is most alike with Nox2 and Nox5 mostly unrelated [243].

Therefore, the NADPH oxidase (Nox) family (Fig. 4.15) is formed from 7 catalytic subunits termed Nox1-5 and Duox1 and Duox2 (for Dual Oxidase), regulatory subunits p22phox, p47phox or Noxo1, p67phox or Noxa1, p40phox. Further, the Nox1, 2, 4 and 5 enzymes are existent in normal cardiovascular tissues, and trigger the progression of cardiovascular disease. Nox enzymes are located in VSMCs, ECs, adventitial fibroblasts, macrophages, CMs and fibroblasts, plus adipocytes and stem cells. They are associated with hypertension, atherosclerosis, HF, ischemia reperfusion injury and CR, but upregulation can be physiologically beneficial such as in angiogenesis [243, 244, 255, 256]. The acutely upregulation of cardiovascular NADPH oxidase activity by a large various patho-physiological stimuli comprise [243] (a) G-protein coupled receptor agonists such as Ang II and ET-1; (b) growth factors such as VEGF, thrombin, PDGF, and EGF; (c) cytokines such as TNF-β, IL-1 and TGF-β; (d) metabolic factors such as elevated glucose, insulin, free fatty acids, and advanced glycation end products (AGE); (e) oxidized LDL, lysophosphatidylcholine, and hypercholesterolemia; (f) mechanical forces such as oscillatory shear stress; and (g) ischemia-related stimuli such as nutrient deprivation, membrane depolarization, flow cessation, hypoxia–reoxygenation, and ischemia [243].

Fig. 4.15
figure 15

Structure of NADPH oxidase in the heart. NADPH oxidase complex is composed of two major components. Plasma membrane spanning cytochrome b558 composed of p22phox and a Nox subunit (gp91phox (Nox2), Nox4) and cytosolic components composed of four regulatory subunits (p47phox, p67 phox, p40 phox and Rac1). The low molecular weight G protein rac1 participates in assembly of the active complex. Upon activation, cytosolic components interact with cytochrome b558 to form an active NADPH oxidase enzyme complex, resulting in release of ·O2−. The primary Nox subunit isoforms in cardiac cells are Nox2 and Nox4. Nox4 oxidase localizes intracellular organelles around the nucleus. The activity of Nox4 results in the direct release of hydrogen peroxide (H2O2) in mitochondria. The mechanisms underlying the generation of hydrogen peroxide by Nox4 oxidase are yet to be fully characterized. (From Kayama et al. [255]. It is an open access article)

Nox2 and Nox4 are the mainly isoform exhibited in CMs. Triggered Nox2 is mainly exhibited at the plasma membrane [244]. According to data, Nox derived ROS are implied in CM apoptosis. Pro-apoptotic signaling pathway and generation of CaMKII in pro-apoptotic signaling pathway are triggered both by Ca2+, by Nox2-derived ROS, and downstream of Ang II [257]. Norepinephrine, aldosterone, and doxorubicin are also reported to promote CM apoptosis through the activation of Nox2 [258,259,260]. Contrary to Nox2 function in Ang II-induced cardiac hypertrophy, Nox2 is not implied in cardiac hypertrophy induced by pressure overload (Fig. 4.16) [261]. The major agonists and stimuli of Nox2 activation in CMs and ECs comprise G-protein coupled receptor agonists (GPCRs) such as Ang II and ET-1, growth factors, cytokines (TNF-α), mechanical forces, metabolic factors (glucose, insulin), glycated proteins [262], and oxidized low-density lipoprotein (ox-LDL) (Fig. 4.17) [261, 263, 264]. To sum up, evidence supports different functions for Nox2 and Nox4 in hypertrophic reaction to pressure overload [243]. Important redox-sensitive downstream signaling pathways in the heart that can be affected by NADPH oxidase activation such as RAS, the MAPKs (p38MAPK, ERK1/2, JNK), c-src, p90RSK, the PI3 kinase (PI3K)/Akt pathway, AP-1, NF-ĸB, HIF-1, and others [243].

Fig. 4.16
figure 16

Representative immunofluorescence micrographs of human heart sections labeled for the nicotinamide adenine dinucleotide 3-phosphate (reduced form) oxidase subunit gp91phox. Panels a, c, e, and g show nonfailing heart tissue and panels b, d, f, and h show end-stage failing tissue. Transverse (a, b) and longitudinal (c, d) sections labeled for gp91phox show increased labeling in end-stage heart failure. Labeling for alpha-actinin (e, f) shows a typical intracellular pattern of myocyte costamer and intercalated disc labeling. Panels g and h show suprerposition of gp91phox and alpha-actinin labeling. All scale bars = 20 μm. (From Heymes et al. [244] with permission.)

Fig. 4.17
figure 17

Schematic illustrating involvement of Nox2 NADPH oxidase in the cardiac response to activation of the renin angiotensin aldosterone system (RAAS) or to chronic pressure overload. Hypertrophy in response to short-term RAAS activation is dependent upon Nox2, whereas the hypertrophic response to pressure overload is not. However, Nox2 is essential for the development of interstitial fibrosis in response to either stimulus. (From Murdoch et al. [261] with permission.)

In case of RV, metabolic and ischemic modifications typical to RV remodelling are also correlated with accumulation of ROS [21, 265] (Fig. 4.18). The presence of ROS activates the cellular and molecular modifications with decrease of contractile function, lacking of energy production and fibrosis. Alteration of SM channels by oxidative stress produces damaging of RyR2 activation and decrease of sarco/endoplasmic reticulum Ca2+-ATPase (SERCA) activity, as a result appears temporary malfunction of myocyte Ca2+ and contractile dysfunction [21, 92]. Additionally, increased ROS amounts cause conversion of nitrotyrosine rests in TIMPs and discharge active MMPs with CR and fibrosis [21, 266]. To date, vivo studies with histological examination of collagen content in RV samples from pulmonary artery showed a significant raise of ROS, important collagen deposition with high levels of MMP-2, MMP-9 and MMP-13 and diminished TIMP-4 protein amounts. Additionally, ROS are second messengers within CMs for numerous signalling molecules (ATII, TGFβ1, TNFα and ET-1) to generate hypertrophic pathways including MAPKs, PKC and Src [21, 62]. Taken together, raised amounts of ROS can damage cellular, molecular and structural components with CR and failure. It is important to underline that malondialdehyde levels represent an indirect index of oxidative stress and are notably elevated in the RV in comparison with the LV. To sum up, these features support a decreased resistance of RV in oxidative stress being a contributor in the development of HF [267].

Fig. 4.18
figure 18

ROS-induced intracellular changes in cardiomyocyte . The increased intracellular ROS levels occurring in RV pressure overload affect several cardiomyocytes functions. ROS can stimulate pro-hypertrophic pathways by targeting key molecules in this process, such as MAPK, PKC and Src proteins. The redox-mediated activation of target transcription factors (HIF-1α, cMyc and FOXO1) might be responsible for the abnormal PKD activation, which inhibits mitochondrial oxidative metabolism, leading to mitochondrial dysfunction. Sustained ROS levels cause mPTP opening and mitochondrial membrane depolarisation. As a consequence, more ROS are produced and cytochrome c is release from mitochondria causing cell apoptosis. HIF-1α activation also decreases the activity of the O2-sensitive Kv channel (Kv1.5), resulting into membrane depolarisation and elevation of cytosolic Ca2+. The surplus of cytosolic Ca2+, in addition to the excessive Ca2+ released from the sarcoplasmic reticulum, as a consequence ROS-mediated RyR2 channel activation and SERCA inhibition, contributes to myocytes contractile dysfunction. ROS are also responsible for the MMPs/TIMPs imbalance that drives ECM remodelling and fibrosis. Antioxidant compounds, like Folic acid or EUK-134, by scavenging the ROS in excess, can help restore the impaired cardiomyocyte function. Furthermore, DCA can restore ROS production and mitochondrial membrane potential by inhibiting PDK and thereby improving glucose oxidation. “Upwards arrow” indicates increase in levels; “Lowerwards arrow” indicates decrease in level. ROS reactive oxygen species, PCK protein kinase C, MAPK mitogen-activated protein kinase, mPTP mitochondrial permeability transition pore, PDK pyruvate dehydrogenase kinase, HIF hypoxia-inducible factor, FOXO1 Forkhead box protein O1, cMyc v-myc avin myelocytomastosis viral oncogene homologue, RyR2 ryanodine receptor 2, Kv 1.5 potassium voltage channel, SR sarcoplasmic reticulum, SERCA sarcoplasmic reticulum Ca2+-ATPase, MMP matrix metalloproteinases, TIMP tissue inhibitor metalloproteinases, ECM extracellular matrix, DCA dichloroacetate, PKD protein kinase D. (From Iacobazzi et al. [21]. It is an open access article)

Peripheral Monocytosis is a sign of monocyte and macrophage infiltration of the necrotic myocardium which arises two to three days after an acute MI. Likewise, a higher peak monocyte level is related with a larger LV end-diastolic volume and inferior LVEF. It was shown that a peak monocyte count ≥900/μL independently predicts HF, LV aneurysm formation, and cardiac events [219]. It should be stressed that monocytes have the capacity to generate and discharge inhibitory mediators of inflammation such as IL-10 and TGF-β [268]. There are varied monocytes with different functions in inflammatory response showed in humans such as CD16-monocytes that exhibit important amounts of CCR2 with pro-inflammatory properties same to murine Ly6Chi cells [268]. Further, inhibition of inflammatory signal pathways is correlated with Ly6Clo/CX3CR1hi monocytes entrapment that generates angiogenic mediators with infarct healing. On the other hand, in patients with ST elevation MI, CD14+/CD16-cells have an early peak and are negatively correlated with heart recovery [269].

6 Factors Influencing Cardiac Remodelling

6.1 Myocardial Infarction

It is the most frequent condition in which CR comes about. Taken together, heart ischemia leads to ‘necrotic cell death’. Further, the post-MI evolution implies apoptosis, inflammation, ECM remodelling, fibrotic scar formation, proliferation and differentiation of MyoFb, angiogenesis, and scar maturation [21]. All these reactions are determined to cause healing on short term, but they produce evolution to HF on long time. Therefore, after MI occurrence, the poor evolution continues with additional CR, hypertrophy, dilation, and systolic dysfunction [21, 73].

A number of innate immune pathways are triggered in MI [149]. It appears the production of “damage-associated molecular patterns (DAMP) ” by necrotic cells that further trigger membrane-bound “Toll-Like Receptors” (TLRs) [270, 271]. Also, among others innate immune pathways such as the “High mobility group box 1” (HMGB1), the “receptor for advanced glycation end-products” (RAGE) [21, 272] and the complement system are also triggered in the onset of inflammation after MI [21]. As a result, ROS are produced at ischemic injury with further activation of inflammatory signals pathways and myocardial dysfunction [149]. All triggered “innate immune pathways ” set off Nuclear Factor NF-κB with further initiation of inflammatory cytokines and chemokines [273]. As already described, pro-inflammatory cytokines significantly modulate the inflammatory reaction to cardiac ischemic injury. IL-1 triggers chemokines production in MI with entrapment of leukocytes [274]. An inactive precursor named pro-IL-1β generates active IL-1β by the converting enzyme caspase-1. Further, caspase-1 function is strongly controlled in multiprotein complexes named “inflammasomes ”, which further monitor production of IL-1β [275]. In MI, “inflammasome” initiation is restricted only in leukocytes and CFs with IL-1-mediated inflammatory cell infiltration and cytokine production [276]. ROS production and K+ efflux have a significant function in inflammasome triggering from CFs. Importantly, chemokines activation is a significant finding of post-MI inflammation [277]. The activation of chemokines receptors from leukocytes in MI exhibit a chemokines profile that controls the composition of the leucocytes infiltrate. Therefore, neutrophils are triggered firstly in MI followed by monocytes and lymphocytes. Apoptotic neutrophils as negative mediators of inflammation are exposed in Fig. 4.19 [149].

Fig. 4.19
figure 19

The role of neutrophil clearance in suppression of the inflammatory response. Abundant neutrophils infiltrate the infarcted myocardium. Neutrophils are short-lived cells that undergo apoptosis; dying neutrophils may contribute to repression of the post-infarction inflammatory response through several distinct mechanisms. First apoptotic neutrophils may release lactoferrin, an inhibitor of granulocyte transmigration. Second, during clearance of apoptotic neutrophils, macrophages secrete large amounts of anti-inflammatory and proresolving mediators including IL-10, TGF-β and lipoxins. Third, expression of decoy cytokine receptors by neutrophils may promote cytokine scavenging. Increased expression of chemokine receptors (such as CCR5) in apoptotic neutrophils may serve as a molecular trap for chemokines terminating their action. (From Frangogiannis [149] with permission)

The activation of neutrophils in the MI triggers apoptosis. Shortly, they are eliminated in MI by macrophages which activate powerful inhibitory pathways. Mediators such as TNF-α and IL-1β can maintain activated neutrophils in MI [278]. On the other hand, within 3–7 days after MI, the neutrophils undertake apoptosis [279].

6.2 Changes in Hemodynamic Load

In case of patients with anterior MI, the early LV dilation may be increased, as well as ventricular hypertrophy turns up to be a late and restricted modification during the first year [41]. Generally, the outcome of ongoing CR with ventricular dilation and abnormal ventricular hypertrophy causes a significant growing in total LV wall tension [41, 280]. As it will become evident, triggering of wall stress can activate further an amount of mechanisms that in the absence of any efficient therapy may cause further CR with progressive HF [50, 281].

6.3 Blood Pressure

Correspondingly, high blood pressure (BP) triggers structural modifications in the LVH with interstitial alterations, which further may produce diastolic dysfunction with HF. Additionally the functional effect of pressure overload hypertrophy may be determined by the features of the CR process. For example, if remodelling is eccentric with LV dilatation with normal relative wall thickness and raised wall stress [282], HF by a functional damage was described. On the other hand, HF does not occur in animals with concentric CR defined as normal chamber volume, raised relative wall thickness and normal wall stress. Previous hypertension may be related with extensive damaging CR and progression of HF after MI. This finding was shown by Richards et al. in 1093 patients, where 68% experienced serial neurohormonal sampling and assessment of LV function one to four days and three to 5 months post-MI [283]. In this study, in comparison with normotensives, hypertensive patients had significantly higher plasma levels of neurohormones at serial sampling with a significantly enhanced raise in LV volumes by remodelling at five months. Conversely, only normotensive patients had a recovery in LV ejection fraction at five months. Also, previous diagnosis of hypertension was related with a greater risk of HF necessitating hospitalization at a mean follow-up of two years (12.4 versus 5.5% in normotensives) [283]. Moreover, Cingolani et al. discovered that TSP-4 from CMs, adjusts cardiac contraction function to acute stress and it has a major role in chronic CR and HF [21, 284].

6.4 Neurohormonal Activation

Progressing HF is connected with an initially compensatory neurohumoral activation that may be a factor to the development of the structural defects. Both the sympathetic system and the renin–angiotensin–aldosterone system (RAAS) are implied in CR. Triggering of both systems turns on intracellular signaling pathways that increase the production of protein in CMs and CFs, with hypertrophy, fibrosis, switching on of growth factors and MMPs [285,286,287]. Moreover, it appears hemodynamic overload by vasoconstriction and water retention, raise of oxidative stress activity with direct cytotoxic effect, and apoptosis [285,286,287]. Therefore, the inhibition of these systems can has a major therapeutic role in attenuation or prevention of CR. Unfortunately, elevated plasma norepinephrine, renin activity, and antidiuretic hormone levels [288, 289] are indicators for poor survival in these patients [290]. Even if, neurohumoral activation is firstly adjustable, it is damaging over the long term by pathologic remodelling, especially in case of Ang II and norepinephrine [291]. The studies data are most convincing for the activation of the RAAS. Also, the plasma BNP concentrations are raised in progressive HF and interrelated with prognosis [292]. In spite of this, the release of BNP from myocytes in HF may defend against pathologic remodelling [293].

The RAAS has a significant function in the control of BP and electrolyte equilibrium. Within RAAS, Ang II produces triggering of sympathetic nervous system with vasoconstriction, sodium and water retention, and anorexia [294]. The damaging effects of RAAS in cardiovascular tissues cause CR by local triggering of the RAAS with autocrine and paracrine mechanisms [295,296,297]. Mainly, the pathophysiological effects of Ang II in the cardiovascular system are controlled by a member of the GPCR family termed the 7 transmembrane (TM7) spanning AT1 receptor [296, 298, 299]. According to recent data, the mechanical stress together with systemically and locally Ang II cause by the triggering of AT1 receptor, cardiac hypertrophy [295,296,297, 300]. It seems that studies with the AT1 receptor blockers (ARBs) as candesartan showed that switching off of triggered AT1 receptor by mechanical stress, notably reduced hypertrophic reaction in cultured CMs [301,302,303]. Therefore, mechanical stress causes cardiac hypertrophy in vivo by initiation of the AT1 receptor with no correlation of Ang II [295, 296].

Mechanical stretch and Ang II by attachment to the AT1 receptor causes to its structure to switch on with occurrence of Cys residues inside the ligand-binding pocket. Further, if mechanical stress continues, TM7 undertakes a counter clockwise rotation with a modification in the ligand-binding pocket [304]. It is not determined exactly by current studies the mechanisms by which mechanical stress activates the AT1 receptor perceives its structure change, preparing for dissimilar initiation of particular intracellular signaling mediators [303].

6.5 Role of Angiotensin II

Significance of angiotensin II (Ang II) in pathologic CR is demonstrated by data of large trials in humans that have been shown that angiotensin-converting-enzyme inhibitors (ACEI) increase survival in HF by decrease or even reverse of some parameters of CR [305, 306]. Shortly, Ang II is produced and has locally and systemically effects. So that, mechanical stretch directly boost Ang II release from CMs [301]. Also, Ang II seems to sustain directly CR. Previous studies have been showed that human CFs cultured from cardiomyopathic and ischemic hearts have on CMs the expression of AT1 receptors [307, 308]. In fact, these CFs reply to Ang II with raise of collagen production by activation of AT1 receptor [309,310,311]. Despite the fact that Ang II is produced locally or systemically, it may directly support CR. In fact, these CFs may reply to Ang II with AT1 receptor-mediated collagen synthesis [309,310,311]. On the other hand, Ang II acts via the AT1 receptor with boosting of protein synthesis and results in hypertrophy of CMs [309]. Both ACEI and Ang II receptor antagonists can reverse remodelling in HF [212].

Aldosterone secretion is increased by Ang II, and also may be a factor in CR. The heart contains mineralocorticoid receptors and takes out aldosterone after a MI, supporting the post MI remodelling [312]. Moreover, the secondary hyperaldosteronism commonly seen in patients with HF may participate to cardiac hypertrophy and fibrosis [313, 314]. It should be stressed that the benefit effects connected with spironolactone or eplerenone, which both link the mineralocorticoid receptor may result with diminished fibrosis [315].

6.6 Energy Metabolism and Cardiac Remodelling

Ischemia, pressure and volume overload are forms of stress that activates human heart to adapt its metabolic function to use glucose instead of the free fatty acids [316]. It seems that free fatty acids provide the highest quantity of ATP to human heart [317]. On the other hand, the glucose metabolism needs a reduced amount of oxygen consuming for same quantity of ATP synthesis, being the most effective alternative in highest metabolic states such as short-term of severe cardiac stress [318]. As a rule, in normal heart, free fatty acids are the main energy substance representing about 60–90% of energy reserves. Both free fatty acids and glucose metabolites undergo β-oxidation and glycolysis in the citric acid cycle, resulting in FADH2 and NADH. Finally, the obtained energy is accumulated and carried as phosphocreatine (Fig. 4.20) [319].

Fig. 4.20
figure 20

Schematic representation of classic pathways of cardiac metabolism . Substrates are transported across the extracellular membrane into the cytosol and are metabolized in various ways. For oxidation, the respective metabolic intermediates (e.g., pyruvate or acyl-CoA) are transported across the inner mitochondrial membrane by specific transport systems. Once inside the mitochondrion, substrates are oxidized or carboxylated (anaplerosis) and fed into the Krebs cycle for the generation of reducing equivalents (NADH2 and FADH) and GTP. The reducing equivalents are used by the electron transport chain to generate a proton gradient, which in turn is used for the production of ATP. This principal functionality can be affected in various ways during HF thereby limiting ATP production or affecting cellular function in other ways (see text and further Figures for details). IMS mitochondrial intermembrane space, GLUT glucose transporter, FAT fatty acid transporter, MPC mitochondrial pyruvate transporter. (Illustration Credit: Ben Smith). (From Doenst et al. [319] with permission)

During stress is stopped the normal inhibition of glucose oxidation by free fatty acids [303]. It seems that the nuclear receptor peroxisome proliferator-activated receptor-α (PPARα) is a significant contributory factor that changes from fatty acid metabolism to glucose metabolism [316, 320]. Further, Karbowska et al. showed on ventricular biopsies from 5 patients a 54% fall of PPARα protein levels in end-stage HF in comparison with controls [316, 321]. Therefore, CR implies cardiac dysfunction with energy loss due to the disproportion from the oxygen reserve and use, with a diminished free fatty acids oxidation and raised glucose oxidation [316]. In addition, β-oxidation fall leads to deposit of triglycerides and lipotoxicity, mitochondrial dysfunction [316]. Altogether, these modifications causes for myocardial proteins further low levels of energy reserves with oxidative stress and ROS, with their sides effects (Fig. 4.21) [316, 321,322,323,324].

Fig. 4.21
figure 21

Overview of metabolic remodeling and proposed mechanisms linking it to other processes in the progression to HF. Metabolic pathways are blue. Bold lines indicate pathways/processes that are increased or dominant. Thin lines represent pathways/ processes that are decreased. The question marks imply unknown causes/ effects. In general, metabolic remodeling in cardiac hypertrophy and failure is characterized by a shift away from energy production to activation of biosynthetic pathways required for structural remodeling processes such as ventricular hypertrophy and fibrosis. Particularly, fatty acid oxidation is decreased and may not be sufficiently compensated given the lack of increase in glucose oxidation. These alterations and further mitochondrial defects result in ATP depletion. Instead of being oxidized, pyruvate may be preferentially used for anaplerosis to maintain Krebs cycle moieties, which might be increasingly channeled into protein synthesis. Hypertrophic mediators such as MAPKs and NFAT are activated as a result of increased mitochondrial ROS and flux through the HBP, respectively. Overproduction of mitochondrial ROS causes oxidative damage. Although the flux through the PPP is increased, anti-oxidative defense might be inadequate due to the consumption of NADPH by the anaplerotic malic enzyme. Mitochondrial damage and ATP depletion may stimulate autophagy. Increased activity of autophagy and the UPS may contribute to hypertrophy by providing amino acids and other metabolites. Increase in mitophagy may trigger myocardial inflammation by releasing mitochondrial DNA. H hexosamine biosynthetic pathway (HBP), P pentose phosphate pathway (PPP), G glycolysis, A anaplerosis, O oxidation, ETC electron transport chain, ROS reactive oxygen species, UPS ubiquitin-proteasome system. (From Doenst et al. [319] with permission)

In case of RV, the metabolism data is from the LV studies. As already described, the RV has smaller afterload than the LV due the decreased pulmonary vascular resistance [316]. Even if, RV and LV have similar stroke volumes, the RV has near 25% from the stroke volume of the LV because of the low PVR [316, 325]. Extensive “transcriptional, translational and energetic” disturbances to physiologic and pathophysiologic stress occur in RV. The conversion from pressure overload and volume overload to cardiac hypertrophy and later RVF is correlated with the switch from free fatty acids to glucose metabolism for ATP production. As a result, the RVF is an energy-deprived state with deficient ATP amounts. For this reason, PET using specific radioactive tracers provides a complete description of RV metabolism [316]. These metabolic differences might corelate with the dissimilarities in ventricular wall stress and intraventricular pressure dynamics (Fig. 4.22) [316].

Fig. 4.22
figure 22

Right ventricular metabolism . (From Altin et al. [316] with permission)

To sum up, metabolic alteration in the hypertrophied RV imitates metabolic alteration of the hypertrophied and failing LV [316]. Nagaya et al. [326] studied 21 patients with RVH due to PH by using magnetic resonance spectroscopic imaging (MRSI) that associates cardiac structure with metabolic function. They found important RV contraction dysfunction in patients with altered myocardial free fatty acid metabolism. Further, in a study of 16 patients with idiopathic PAH, Bokhari et al. proved that PET imaging is for determining myocardial glucose assimilation and use [316, 327]. They demonstrated that RV glucose usage is associated with hemodynamic parameters such as mean PA pressure, doubtlessly implying that RV dysfunction is switched on myocardial glucose metabolism and being a sign of RV dysfunction. Can et al. [328] have been demonstrated same features on 23 patients with PAH and 16 healthy controls evaluated by PET. Their results established that raised fludeoxyglucose (18F) increase in the RV myocardium were connected with raised RV loading conditions and with the existence of elevated pulmonary artery pressures but not with their stage [316, 328].

Also, MRSI for the study of myocardial triglyceride load demonstrated an accurate statistically significance correlation with triglycerides from RV biopsy [316, 329]. Currently, no other study tried to measure in the RV the lipid transitional products.

It is not determined if cardiac metabolic alterations maintain during the development of RVH, that is characterized by important decrease of CO, increase of RV filling pressure and raised fibrosis [21, 330]. Consequently, RVH is correlated with increased mitochondrial ROS, which downregulates HIF1α and triggers p53 pathways, in the end with dysregulated pyruvate dehydrogenase kinase (PDK) and diminished glucose uptake [21, 331]. Raised PDK expression is a frequent feature in RVH during glucose oxidation, as a result there is a decrease in mitochondrial respiration [21, 330]. A number of clinical trials directed on molecular dysfunction in RVH and RV failure are undergoing, even if the precise outcomes for pharmacologic involvements on RV abnormal metabolism are not determined [316].

6.7 Electrical Remodelling in Cardiac Remodelling

As already been described, the functional, structural, and electrical modifications of CMs to stress described by hypertrophy, HF, and ischemia define CR, which disposes to raised occurrence of ventricular ectopy and arrhythmias. “Arrhythmia-induced changes in the electrophysiological properties of heart tissue which predisposes to an increased frequency of ventricular ectopy and arrhythmias are referred to as electrical remodelling” [332].

Specifically, Na+-Ca2+ exchanger (NCX) mediates intracellular Ca2+ concentration and its activity is controlled by intracellular concentrations of Ca2+, Na+, ATP, pH, and phosphorylation of NCX, all being modified in HF [332]. Therefore, HF with contractile dysfunction and arrhythmogenesis may be explicated by amplified NCX and diminished SERCa2+a function, consequently with raised Ca2+ discharge from CMs and delayed afterdepolarizations [332,333,334]. Also, during cardiac diastole dysfunction, NCX sustains Ca2+ transportation from intracellular space and diminishes SERCa2+a function [332, 333]. It has to be underlined, that NCX activity is mediated in cardiac hypertrophy, function sustained by diminished NCX activity but with raised NCX protein and transcript levels [332]. It seems that calcineurin inhibition weakens the boost of NCX1 transcript and protein levels correlated with pressure overload, advocating that calcineurin is vital in the adjustment of NCX1 transcript synthesis and degeneration [332, 335].

Also, there is a diminished expression and function of Na+-K+ATPase in HF [332, 336] that make susceptible cardiac tissue to arrhythmias by raised action potential duration, and increased depolarizing current, and extracellular K+ [332]. According to experimental cardiomyopathic studies, ETA receptor blockers demonstrated to stop electrical remodelling and ventricular arrhythmias by diminishing K+ and Ca2+ current expression, rising QT interval and action potential duration [332, 337].

Electrical anisotropy caused by myocardial fibrosis and modifications of intracellular Ca2+ could produce the electrophysiological remodelling and arrhythmias from hypertrophy [332]. Further, modification of the collagen amount, type, and cross-linking is correlated with myocardial fibrosis and CR with electrophysiological abnormalities [338].

RVH, dilation, and septal displacement also create RV dyssynchronous motion [339,340,341] and dyssynchronous RV-LV contraction [341,342,343]. Delayed RV lateral wall contraction and interventricular dyssynchrony in PAH are not related to QRS duration or abnormal electric activation such as left bundle-branch block but rather to RV wall stress, septal shift, LV end-diastolic volume, and stroke volume [342, 343]. These ventricular-ventricular interactions almost certainly increase the ratio of systolic to diastolic duration because interventricular dyssynchrony is related to lengthening of the RV contraction [343].

To sum up, existing data supports that myocardial hypertrophy form determines the electrical CR and the reverse of it [332, 344]. Nevertheless, in pressure-overload states with reverse of hypertrophy is correlated with the reverse of the electrical remodelling [345,346,347]. However, the dissimilarity between the reverse of electrical remodelling in pressure-versus volume-overload states is unknowable [332]. Both structure and electrical CR should be regarded as independent clinical disorders [332]. The assessment of risk factors for arrhythmias has a significant function in the regress of hypertrophy and the electrical CR [347].

6.8 Coronary Vascular Remodelling in Cardiac Remodelling

Coronary vascular remodelling causes adjustable reactions such as the rapid adaptation of vessel diameter by modifications in smooth muscle tone, changes in vessel diameter structure, adding or elimination of vessels by “angiogenesis (sprouting/splitting)”, or “vascular pruning” [348]. It important to underline that physiological vascular adaptation keeps an appropriate perfusion, but vascular maladaptation takes place in disorders such as hypertension [348]. Also, regulatory mechanisms in larger vessels are different from microcirculation that has a vital role in physiological vascular adaptation and pathological states [348]. Generally speaking, growing of size and number of microvessels during exercise or involution with microvascular remodelling because of constant decline of physical activity appear [349]. Reduction of epicardial arteries with hemodynamic- and metabolic modifications causes process of collateralization or arteriogenesis defined by “structural enlargement of arteriolar vessels and arterio-arterial anastomoses” [348, 350].

As a result, constant chronic remodelling of coronary vessels leads to over-prolonged modifications of vessels diameter with or without shifts in wall mass (Fig. 4.23) [348, 352, 353]. Therefore, coronary vessels adjust to mechanical stimuli, such as fluid shear stress acting on ECs, circumferential wall stress and metabolic signals [348, 354, 355].

Fig. 4.23
figure 23

Mechanisms and models of vascular remodelling . Left: Signals for vascular adaptation include wall shear stress at the endothelial surface (t), circumferential wall stress (s), and metabolic signals. Metabolic signals maybe elicited by low oxygen availability and act as vasodilators and stimulate vascular growth, or they could be vasoconstricting mediators produced at high oxygen partial pressures but in decreasing amounts with decreasing pO2. Metabolic substances are convected downstream, but elicit also a signal that is conducted upstream within the vessel wall. Middle: Vascular responses elicited by these stimuli comprise changes of diameter and wall mass. Right: An integrated model [351] which connects the local conditions (pressure, flow, and metabolic state) with derived stimuli (σ, τ, and metabolic stimuli) and the vascular changes in vessel diameter or wall mass. Lines indicate biological reactions (solid) and physical relations (dashed). (From Pries et al. [348] with permission)

Essentially, exposure of the LV to afterload stress causes firstly the development of new capillaries or angiogenesis, to sustain the raised blood flow of hypertrophied CMs. In case of angionesis are implied raised production of the proangiogenic factors hypoxia HIF1α and VEGF. If the LV starts to fail, the capillary density starts to decrease. On the other hand, the capillary density of the RV reduces with the beginning of pressure overload upsurge. RVF is a frequent side effect of chronic RV pressure overload with progression to RV ischemia. Specifically, increased pulmonary arterial pressure boosts RV wall tension and oxygen demand in correlation with alteration of coronary blood flow. For instance, Eisenmenger’s syndrome produces a level of RV afterload same to idiopathic PAH, only that the survival is longer with latent overt RVF [356]. In addition, PAH during diagnosis protocol have different degrees of RVF even if the RV afterload is highly developed.

7 Reverse Cardiac Remodelling

CR can be reversed with maximized therapy that initiates ongoing recovery of cardiac function and thus enhances prognosis of patients [357]. Even if, reverse CR could arise unexpectedly in heart pathologies, it is more frequently seen as reaction to medical, device-based, or surgical therapies, such as beta-blockers, cardiac resynchronization therapy (CRT), revascularization and valve surgery [358]. The various cardiac pathologies with noticed reverse CR prove that myocardial remodelling is bidirectional and takes place no matter of the myocardial disease aetiology, length, and severity (Fig. 4.24) [358]. Moreover, prognosis is improved in patients with reverted heart dysfunction, for that reason reversal of CR should be the most important treatment aim. Therefore, effective treatment should reverse cardiac remodelling [359].

Fig. 4.24
figure 24

Myocardial recovery in clinical settings. The segments of the outmost ring highlight pathophysiological processes implicated by reverse remodeling in particular clinical settings that comprise the middle ring. Abbreviations: ALM acute lymphocytic myocarditis, CPAP continuous positive airway pressure, RAAS renin–angiotensin–aldosterone system, LVAD left ventricular assist device, MVR mitral valve repair/replacement, AVR aortic valve replacement, CSD cardiac support device, CRT cardiac resynchronization therapy. (From Hellawell et al. [358] with permission)

Both ACEI and Ang II receptor antagonists, also known as angiotensin receptor blockers (ARBs) have been utilized to prevent CR. Preventing raised RAAS stimulation that may limit subsequent maladaptive cardiac remodelling. For instance, RV samples taken from control patients showed higher ventricular weight with raised collagen and foetal contractile protein genes, and diminished SR Ca2+-ATPase. Moreover, ACEI exert antioxidant effects by inhibiting the transcription factor NF-kB that controls the synthesis of various genes associated with inflammatory response, such as cytokines, chemokines, growth factors, and cell adhesion molecules [360,361,362,363]. The complexity of the intramyocardial mechanisms involved in CR should also take into account endothelial damage, on which PDE5 inhibition acts positively, as recently demonstrated in a meta-analysis of type 2 diabetic cardiovascular patients [364]. Also, clinically relevant evidence suggests that PDE5 inhibition has favourable direct myocardial effects via cGMP and cAMP activities that may counterbalance hypertrophic and proapoptotic signaling, including adrenergic stimulation [365].

7.1 Cardiac Regenerative Medicine

Production of “induced cardiac-like myocytes” (iCLMs) shows all the signs of a new future successful method to regenerate damaged CMs [366]. Replacing lost CMs by injecting cardiac progenitors, cardiospheres, or CMs derived from “induced pluripotent stem cells” (iPSCs) and/or “embryonic stem cells” (ESCs) has been researched intensively [367]. Importantly, miRNAs are important for stem cell differentiation, as well as indirect and direct reprogramming to multiple lineages [368,369,370,371]. To sum up, miRNA based therapy can be used to promote CMs proliferation, reprogram directly fibroblasts to CMs or indirectly to iPSc as well as driving the differentiation of iPSCs, ESCs or CPCs to CMs (Fig. 4.25) [367]. Another new treatment choice is to release cells in the damaged myocardium. According to evidence, various cell categories have been utilized for heart regeneration, such as ESCs, CMs obtained from iPSCs, mesenchymal stem cells (MSCs), bone marrow MSCs, cardiac stem cells, cardiac progenitor cells, skeletal myoblasts, ECs, adipose tissue-derived stem cells (ATDSCs), and CMs [372]. Nevertheless, studies have still unremarkable outcomes.

Fig. 4.25
figure 25

miRNAs and reprogramming. miRNAs promote the generation of cardiomyocytes via a number of mechanisms. Fibroblasts can be reprogrammed into cardiomyocytes by miRNAs directly or through an intermediate iPSC state. miRNAs also promote cardiac progenitor cell (CPC) and embryonic stem cell (ESC) cardiac differentiation. miRNAs can promote or inhibit cardiomycyte proliferation. (From Hodgkinson et al. [367] with permission)

7.2 Device-Based Therapies

Pharmacological treatments that diminish on either side PVR or systemic vascular resistance can reduce the development of fibrosis in the RV and LV, respectively. Likewise, non-pharmacological mechanical decrease of LV load by LV assist devices (LVADs) can attenuate fibrosis in both ventricles [373].

For instance, cardiac resynchronization therapy (CRT) with biventricular (BiV) pacing is an well-known choice therapy in case of patients with overt HF, diminished LV systolic EF, and delayed ventricular conduction with enlarged QRS complex (e.g. electrical dyssynchrony). In fact, Sachse et al. [374] showed that CRT reduces symptoms and mortality in patients affected by dyssynchronous heart failure (DHF) produced by dyssynchronous electrical and mechanical activation of the left and right ventricle. Also, they concluded that CR of electrophysiological properties, hemodynamic and protein expression due to DHF is partially restored by CRT. Specifically, CRT can reverse damages of intracellular structures and function of CMs from HF, with early successful signs of recovery as tubular system structure [374].

Ventricular Assist Devices (VADs) can prevent ongoing CR and produce reverse CR, mainly by decrease of mechanical load of the damaged ventricles [373]. Also, VADs are not same thing with artificial hearts, which are planned for temporary taking up of cardiac function with their latter withdrawal from the patient’s heart [373]. VADs are designed to support the LVAD, the right ventricle (RVAD) or both ventricles (BiVAD). LVAD is the most frequent device used in a damaged heart, but in case of increased pulmonary arterial resistance, RVAD will be added to help with cardiac circulation. According to data, the evaluation of LVAD outcomes looking the reverse of RV remodelling, showed there were no change in CMs size among patients with LVADs and control group [373]. Conversely, LVAD therapy produced diminishing of collagen and TNF-α from RV, supporting that LVAD can reduce RVH by inhibiting the paracrine factors [373]. Further, Barbone et al. studied the involvement of these factors in reverse of CR [375]. Shortly, they studied heart samples from patients who required either LVAD or pharmacological treatment for severe HF [375]. Regardless of LVAD type, the RV volume and CM size enlarged, but isolated RV muscle pieces from inserted LVAD showed a diminished force formation at high pacing rates. Finally, they concluded that reverse RV remodelling after LVAD placing is minimum [375]. In conclusion, LVADs implantation is largely helpful for showing biology of reverse CR: changes of mRNA and microRNA profiles, decrease of apoptosis, diminishing of inflammatory cytokines (e.g., TNF-α), ECM remodelling, regression of action potential lengthening, regression of cardiac myocyte hypertrophy, improved contractility, regression of shape distortions, and improved β-adrenergic responsiveness [358].

Other studies evaluated hearts from end-stage CHF with no VADs, with LVAD or with BiVAD [376]. In comparison with LVAD, BiVAD-supported hearts showed notably diminished right atrium pressures with nearly normal RV end-diastolic pressure-volume interactions. Moreover, the hearts with BiVAD demonstrated normalized RV myocyte diameter and myocardial contraction when isoproterenol perfusion was used. All these modifications were not demonstrated in hearts with LVADs. However, LVADs diminish only RV afterload. To sum up, VADs could induce RV remodelling, but further studies data is necessary [376].

Conclusions

RV remodelling is correlated with functional, cellular and molecular changes [316]. CMs hypertrophy and hyperplasia modify RV geometry, while apoptosis rate and damages of intracellular structures induce further remodelling [316]. RVH and RV dilatation can be correlated with diminished ventricular volume, associated or not with changed hemodynamic status [316]. As a result, physiologic and pathophysiologic stress produces in RV serious transcriptional, translational and energetic modifications [316]. The development of cardiac hypertrophy from pressure and volume overload with failing is correlated with transition from free fatty acids to glucose metabolism for ATP formation [316]. RVF is an energy-depleted condition with lacking ATP levels. Reassuring successful molecular targets are established by ongoing clinical trials studying the molecular alteration from RVH and failure [316]. Ultimately, medical therapy with vasodilators seems to raise both RV stroke volume and CO, while ACEI and ARBs can postpone RVH by diminishing the exhibition of hypertrophy-related genes in the RV. Nonetheless, immunomodulator therapy is correlated with diminishing of RVH and remodelling-related gene expression [316].