Next Article in Journal
Literature Review of Cervical Regeneration after Loop Electrosurgical Excision Procedure, and Study Project (CeVaLEP) Proposal
Next Article in Special Issue
Treatment of Transthyretin Amyloid Cardiomyopathy: The Current Options, the Future, and the Challenges
Previous Article in Journal
Acid-Base Disorders in COVID-19 Patients with Acute Respiratory Distress Syndrome
Previous Article in Special Issue
RBM20-Related Cardiomyopathy: Current Understanding and Future Options
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Genetic Insights into Primary Restrictive Cardiomyopathy

1
Erich and Hanna Klessmann Institute, Heart and Diabetes Center NRW, University Hospital of the Ruhr-University Bochum, Georgstrasse 11, 32545 Bad Oeynhausen, Germany
2
Comprehensive Heart Failure Center (CHFC), Department of Medicine I, University Clinic Würzburg, Am Schwarzenberg 15, 97078 Würzburg, Germany
*
Authors to whom correspondence should be addressed.
J. Clin. Med. 2022, 11(8), 2094; https://doi.org/10.3390/jcm11082094
Submission received: 14 March 2022 / Revised: 5 April 2022 / Accepted: 6 April 2022 / Published: 8 April 2022
(This article belongs to the Special Issue Cardiomyopathies: Current Treatment and Future Options)

Abstract

:
Restrictive cardiomyopathy is a rare cardiac disease causing severe diastolic dysfunction, ventricular stiffness and dilated atria. In consequence, it induces heart failure often with preserved ejection fraction and is associated with a high mortality. Since it is a poor clinical prognosis, patients with restrictive cardiomyopathy frequently require heart transplantation. Genetic as well as non-genetic factors contribute to restrictive cardiomyopathy and a significant portion of cases are of unknown etiology. However, the genetic forms of restrictive cardiomyopathy and the involved molecular pathomechanisms are only partially understood. In this review, we summarize the current knowledge about primary genetic restrictive cardiomyopathy and describe its genetic landscape, which might be of interest for geneticists as well as for cardiologists.

1. Introduction

In clinical practice, cardiomyopathies are divided according to structural and functional criteria into different classes [1,2]. Classification according to their etiology revealed a non-negligible percentage of genetic cases for all structural cardiomyopathies [3]. In comparison to hypertrophic cardiomyopathy (HCM) with an estimated prevalence of 1:500 [4], the prevalence of restrictive cardiomyopathy (RCM) is currently unknown [5]. However, because of the rarity of primary RCM, its genetic background is poorly defined compared with other cardiomyopathies. Beside primary RCM, it can manifest as a part of systemic diseases such as amyloidosis [6], which can also be genetically caused, for example, by mutations in the TTR (transthyretin) gene [7]. In addition, RCM can also be part of different syndromic diseases, e.g., Alström syndrome (MIM, #203800) [8] or Myhre syndrome (MIM, #139210) [9]. In this review, we will focus on the genetic etiology of primary RCM and will summarize the current knowledge of the RCM-associated genes.

2. Clinical Description

RCM is characterized by severely enlarged atria, normal-sized ventricles, with increased myocardial stiffness leading to impaired ventricular filling and diastolic dysfunction (Figure 1). Systolic function and ventricular wall thicknesses are often normal. Patients present with symptoms of left and/or right ventricular heart failure with preserved ejection fraction (HFpEF), atrial fibrillation, ventricular arrhythmias and frequently conduction disorders [10]. The overall prognosis is poor and the 5-year survival rate of adult patients with a confirmed genetic cause was 56% [11]. Specific therapies of non-infiltrative genetic forms do not exist. Non-specific therapies include fluid and sodium restrictions and medical treatment of heart failure with reduction of volume overload as well as anticoagulation and antiarrhythmic therapy. Very often heart transplantation (HTx) is the only option for long-term survival [12].

3. Genetic Landscape of Restrictive Cardiomyopathy

Currently, pathogenic mutations in 19 different genes have been identified in patients with RCM (Table 1 and Figure 2A). Since RCM is a rare cardiomyopathy with an unknown prevalence [13], the genetic landscape is not completely discovered. At present, for several of the known RCM genes, only a single family or even a single index patient has been reported. All known RCM genes are localized on autosomes (Figure 2B) and in most cases, the mutations are inherited in an autosomal dominant mode or appear as de novo mutations. However, there are also some examples for a recessive inheritance pattern [14]. The majority of RCM genes encode for sarcomere, cytoskeleton or Z-disc proteins, e.g., the cardiac troponins, desmin or filamin-C (Figure 2A). Remarkably, there is a significant genetic overlap with other cardiomyopathies especially with HCM and to some extent with dilated cardiomyopathy (DCM), left-ventricular non-compaction cardiomyopathy (LVNC) or arrhythmogenic cardiomyopathy (ACM) (Figure 3). Currently, it is unknown why mutations in the same gene cause different cardiomyopathies. However, additional genetic modifiers as well as diverse environmental factors can be suggested to be contributing to these phenotypical differences. Sometimes, different phenotypes including RCM are even present within the same family [15,16].
Remarkably, there is also a genetic overlap between RCM and myofibrillar myopathy (MFM, MIM, #601419). MFM is a group of genetic muscle diseases characterized by myofibrillar disorganization and abnormal intra-sarcoplasmic protein aggregates [17]. It can affect the skeletal and/or cardiac muscle. Mutations in seven genes cause MFM (DES [18], CRYAB [19], FLNC [20], LMNA [21], BAG3 [22], TTN [23,24], MYL2 [25]) as well as RCM (Table 1). The genetic overlap between both diseases (Figure 3) might indicate a detrimental involvement of pathological cardiac protein aggregates [26].
Table 1. Overview about RCM-associated genes and proteins.
Table 1. Overview about RCM-associated genes and proteins.
GeneCytogenetic LocationEncoded ProteinSubcellular Protein
Localization
First DescriptionReferences
TNNI319q13.42cardiac troponin ISarcomere2003[27]
TNNT21q32.1cardiac troponin TSarcomere2006[28]
DES2q35desminIntermediate filament2006[29]
ACTC115q14cardiac actinSarcomere2008[30]
MYH714q11.2β myosin heavy chain Sarcomere2008[31]
TPM115q22.2tropomyosin 1Sarcomere2011[32]
MYL33p21.31essential myosin light chain 3Sarcomere2011[32]
MYL212q24.11cardiac regulatory myosin light chain Sarcomere2011[32]
MYPN10q21.3myopalladinSarcomere, Z-disc2012[33]
TTN2q31.2titinSarcomere2014[34]
MYBPC311p11.2cardiac myosin binding protein CSarcomere2015[35]
TNNC13p21.1cardiac troponin CSarcomere2016[36]
FLNC7q32.1filamin CIntercalated disc, Z-disc, sarcolemma2016[37]
TMEM87B2q13transmembrane protein 87 BMembrane2016[38]
ACTN21q43α actinin 2Z-disc2016[39]
CRYAB11q23.1αB crystallinIF associated protein, intercalated disc, Z-disc2017[40] 1
LMNA1q22lamin A/CNuclear lamina2018[41]
BAG310q26.11bcl2 associated athanogene 3Cytosol2018[42]
DCBLD23q12.1discoidin cub and lccl domain containing protein 2Membrane2021[43] 2
1 RCM-associated with skeletal myopathy. 2 RCM-associated with atrial fibrillation, tachycardia, developmental delay and dysmorphic features.

3.1. Mutations in Genes Encoding for Sarcomere Proteins

The majority of known RCM-associated mutations are found in ten genes encoding for sarcomere proteins (Figure 2A). These mutations affect the thin and thick filaments as well as titin filaments.

3.1.1. Cardiac Troponins (TNNI3, TNNT2, TNNC1) and Alpha-Tropomyosin (TPM1)

The cardiac troponin complex is composed of three subunits controlling the position of tropomyosin, essential for the regulation of striated muscle contraction and located along the sarcomere thin filament [44]. Disruption of regulatory function due to mutations leads to cardiac dysfunction and cardiomyopathy. Since the early 1990s, cardiac troponins are known as disease genes for HCM [45], however, they expand their disease spectrum to all genetic forms of cardiomyopathies including RCM.
The gene encoding the cardiac isoform of troponin I (TNNI3) is the main target gene for RCM within the thin filaments and the sarcomeres. Almost all mutations are located in the regulatory C-terminal region interacting with actin and the N-terminal domain of TNNC1 (Table 2 and Figure 4). A high proportion of de novo mutations in infants and children with a poor outcome are described. Few mutations are solely reported to cause an RCM phenotype, but most of them are also found in patients with HCM. Studies on skinned fibers by Gomes et al. suggest that TNNI3 mutations increase Ca2+ sensitivity of force development and decrease the ability of TNNI3 to inhibit actomyosin ATPase activity, leading to impaired relaxation properties and diastolic dysfunction [46]. Additionally, it has been shown that mutant alleles, such as p.L144Q, p.R145W and p.R170W, incorporate into the thin filaments to a lower extent compared to wildtype affecting the structural stability of the filaments [47,48]. Overall, it appears that similar mutations can cause a hypertrophic, dilated or restrictive phenotype assuming that genetic modifiers or other environmental factors influence the age of onset and phenotypic expression. A transgenic mouse model (cTNI-193His) corresponding to the human p.R192H mutation mimics the RCM phenotype in mice and suggests that impaired relaxation resulting from Ca2+ hypersensitivity [49] and diastolic dysfunction occurring in a dose-dependent manner and indicating that the dosage of mutant protein may be important for the severity of impaired diastole [50].
In contrast to TNNI3, a restrictive phenotype appears to be less common in the two other troponin genes. Mutations in TNNT2 are mainly reported in rare cases where other cardiomyopathy phenotypes also occur in the same family. Furthermore, two compound heterozygous mutations in the cardiac TNNC1 evolved in a restrictive phenotype in two infants (Table 2) [36]. Kawai et al. developed a knock-in mouse model (TnC-A8V), which mimics the human phenotype of enlarged atria, hyper contractility and diastolic dysfunction. The authors suggest perturbed cross-bridge kinetics by myosin rod hypophosphorylation as a potential novel mechanism [51].
Alpha tropomyosin (encoded by TPM1) is a long, double-stranded, helical coiled-coil protein that is wrapped about the long axis of the actin backbone (Figure 4, red structure) and serves to block the active site on actin, thereby inhibiting actin and myosin from binding under resting conditions. TPM1 and the troponin complex constitute the Ca2+-sensitive switch that regulates the contraction of cardiac muscle fibers. Several missense mutations have been described causing either HCM or DCM [52]. Recently, Dorsch et al. reported a 6-year-old child with severe RCM carrying two TPM1 variants in compound heterozygous state requiring HTx, whereas family members with one of the two variants expressed an HCM-like phenotype [16]. In summary, the one case indicates that TPM1 is a very rare disease gene and the RCM phenotype may only occur in compound heterozygosity.
Table 2. Overview about known RCM-associated thin filament mutations.
Table 2. Overview about known RCM-associated thin filament mutations.
MutationAge of Onset and Clinical FeaturesFamily HistoryMAF 1CommentsReferences
TNNI3
p.D127Yinfant, HF, VADde novo-contractile dysfunctions and effects on thin filament structure[53]
p.L144Qadult, HFunknown- [27]
p.L144Hyoung adults, HFfamilial- [54]
p.R145Wchildren and adults, HFfamilial, autosomal dominant3/280226variant also associated with HCM; Dutch founder mutation; segregation in several families[27,39,55]
p.R145Qchildrenfamilial, far relative HCM-associated with HCM[55]
p.S150Pchild, SCDfamilial-one Chinese family with several affected members[56]
c.549+2delTinfant, died at age 2de novo-predicts splicing defect and truncation[55]
p.D168fsX176child, HF, died at age 28yde novo-protein reduction[57]
p.R170Gchild, HFde novo- [47]
p.R170Winfantde novo-variant also associated with HCM[47,58]
p.R170Qchild, HFde novo-variant also associated with HCM[30,54]
p.A171Tadult, HF, AFunknown- [27]
p.E177fsX209childde novo- [30]
p.K178E6y, HFde novo- [27]
p.K178delchildde novo- [55]
p.D190Hmainly adults, HF, SCDfamilial-named in ClinVar as p.D190G[27]
p.R192Cchildfamilial-carries also mosaicism of p.R145Q; associated also with HCM in far relative[55]
p.R192Hchildren, young adult, HFde novo-independent reports of de novo mutations; variants also associated with HCM[27,59,60]
p.K193Eadults, AF, SCDfamilial-cousin developed HCM[61]
p.I195fsyoung adult, HF, HTxde novo-dominant-negative effect[62]
p.D196Hthree adults, HF, HTxfamilial, homozygous-heterozygous carrier asymptomatic [63]
p.R204Hchildren, HF, HTx, VSD in one casede novo-independent reports of de novo mutations[59,64,65]
TNNT2
p.I89Ntwo adult cases within one familyfamilial0.00002mixed phenotype with HCM and DCM[66]
p.R104Cchildren, young adult, HFfamilial-mixed phenotype with HCM in the family[67]
p.E69delinfant, HF, VADde novo- [28]
p.E146Kchildfamilial0.00003variant also associated with other CMPs[30]
TNNC1
p.A8V; p.D145Etwo infants diedfamilial, compound heterozygous0.00001 0.0001HCM which evolved into RCM[36]
TPM1
p.E62Q;
p.M281T
childfamilial, compound heterozygous-
0.00001
each single variant leads to a HCM like phenotype[16]
ACTC1
p.D313Hchildfamilial-father was diagnosed with DCM[30]
1 MAF = Minor allele frequency according to Genome Aggregation Database (February 2022), https://gnomad.broadinstitute.org (accessed on 13 March 2022). AF = atrial fibrillation, CMPs = cardiomyopathies, DCM = dilated cardiomyopathy, HCM = hypertrophic cardiomyopathy, HF = heart failure, HTx = heart transplantation, RCM = restrictive cardiomyopathy, SCD = sudden cardiac death, VAD = ventricular assist device, VSD = ventricular septal defect.
Figure 4. Schematic molecular structure of the thin filaments in the Ca2+ free state [68] (https://www.rcsb.org/structure/6KN7 (accessed on 13 March 2022)). Actin is shown in light green, tropomyosin is shown in red, cardiac troponin T is shown in blue, troponin C is shown in violet and troponin I is shown in orange. The localizations of the RCM-associated TNNI3 missense mutations are shown in cyan. The majority of RCM-associated TNNI3 missense mutations are localized in the C-terminal part of troponin-I.
Figure 4. Schematic molecular structure of the thin filaments in the Ca2+ free state [68] (https://www.rcsb.org/structure/6KN7 (accessed on 13 March 2022)). Actin is shown in light green, tropomyosin is shown in red, cardiac troponin T is shown in blue, troponin C is shown in violet and troponin I is shown in orange. The localizations of the RCM-associated TNNI3 missense mutations are shown in cyan. The majority of RCM-associated TNNI3 missense mutations are localized in the C-terminal part of troponin-I.
Jcm 11 02094 g004

3.1.2. Cardiac Actin (ACTC1)

Human cardiac α-actin, encoded by ACTC1, is one of the six human actin isoforms. Using fluorescence in situ hybridization technique Ueyama et al. showed that ACTC1 is localized on chromosome 15q14 [69]. Cardiac α-actin is highly conserved between different species and skeletal and cardiac α-actin are co-expressed in cardiomyocytes [70]. As a monomer, actin has a globular structure (G-actin) and polymerize into filaments (F-actin). Actin is the major structural component of the thin filaments (Figure 4, green structure) and is eminent for the contraction cycle and force generation of cardiomyocytes [71].
Kaski et al. described for the first time an RCM causing mutation in ACTC1 (p.D313H) [30]. The father developed DCM and the sister of the index patient showed a mixed RCM/DCM phenotype, but no genetic sequence analysis was performed for both [30]. Functional analysis was not performed in this study. However, ACTC1-p.D313H is localized in the tropomyosin binding region which supports its functional impact. In addition, ACTC1 mutations can cause DCM [72], HCM [73], LVNC [74] and septal defects [75] (Figure 3).

3.1.3. Myosin Heavy and Light Chains (MYH7, MYL2 and MYL3)

The thick filaments of the cardiac and skeletal sarcomere are mainly formed by myosin. Human cardiac myosin is a hexameric protein complex consisting of β myosin heavy chains (encoded my MYH7), two essential light chains (encoded by MYL3) and two regular myosin light chains (encoded by MYL2) [76,77,78]. Myosin proteins consist of a head, neck and tail domain. The head domains interact with the thin filaments and contain the N-terminal globular motor domains [79] performing the power stroke during contraction [80]. The neck region is bound by the myosin light chains [81] and the tail domains build a coiled-coil [82].
In all three myosin genes, mutations have been found in RCM patients (Table 3). For the first time, Karam et al. described in 2008 a de novo mutation in the MYH7 gene (p.P838L) in an infantile patient with RCM [31]. Several further pathogenic MYH7 mutations have been described for RCM (Table 3). The majority of these mutations are missense mutations. Beside RCM, MYH7 mutations are particularly causative for HCM [83] and to a less extent for DCM [84], LVNC [74] and ACM [85].
In 2011, Caleshu et al. reported a female RCM patient carrying MYL2-p.G57E and in addition MYL3-p.E143Khom [32]. The described index patient carrying these myosin light chain variants do not present a family history of cardiomyopathies [32], which might be caused by a recessive inheritance. The mutation MYL3-p.E143Khom was also identified before in the homozygous state in HCM patients [86]. Transgenic mice with the cardiac expression of human MYL3-p.E143K developed an increased ventricular stiffness, cardiac interstitial fibrosis and showed ultrastructural defects of the sarcomeres leading to a restrictive phenotype [87]. MYL2 and MYL3 mutations also cause HCM [88,89] and DCM [90] (Figure 3).
Table 3. Overview about known RCM-associated myosin mutations (MYH7, MYL2, MYL3).
Table 3. Overview about known RCM-associated myosin mutations (MYH7, MYL2, MYL3).
MutationAge of Onset and Clinical FeaturesFamily HistoryMAF 1CommentsReferences
MYH7
p.Y386Cinfant, coronary artery bridgingunknown- [91]
p.R721Kadult, AF, familial-in combination with ABCC9-p.R1186Q [92]
p.G768Radult, AF, unknown- [39]
death at age 42
infant, HTxunknown- [93]
p.R783Hadult, AVB, familial0.00002son has HCM[39]
death at age 54
p.P838Linfantde novo- [31]
p.L840Mchildunknown-in combination with MYBPC3-p.P147L [39]
p.R870Ctwo adults, AFfamilial0.00002myofibrillar disarray, [94]
cardiomyocyte necrosis,
abnormal nuclei morphology
p.I909Madult, AVB, AF, death at age 56unknown- [39]
p.T1188CfsX22adult, in combination with LVHde novo- [95]
MYL2
p.G57Eadultabsent0.000004in combination with MYL3-p.E143Khom[32]
MYL3
MYL3-p.E143Khomadultabsent0.00001in combination with MYL2-p.G57E[32]
1 MAF = Minor allele frequency according to Genome Aggregation Database (February 2022), https://gnomad.broadinstitute.org (accessed on 13 March 2022). AF = atrial fibrillation, AVB = atrioventricular block, HCM = hypertrophic cardiomyopathy, HTx = heart transplantation, LVH = left ventricular hypertrophy, VUS = variant of unknown significance.

3.1.4. Cardiac Myosin Binding Protein C (MYBPC3)

Another main disease gene for HCM and to a minor extent DCM and LVNC is the gene encoding the cardiac myosin binding protein C (MYBPC3). One study by Wu et al. showed that one de novo variant, previously also associated with HCM (p.E334K) and one truncation variant p.Q463X might cause RCM as part of the phenotypic spectrum [35].

3.1.5. Titin (TTN)

Titin is the largest known human protein and represents the third filament system in cardiac and skeletal muscle [96]. Its primary role is maintaining sarcomere organization, generation of passive tension during muscle stretching and modulating contraction. The major cardiac phenotype caused by TTN mutations is DCM, however so far almost exclusively truncation variants are proven to be causative accounting for 30% of affected individuals with DCM [97,98]. Recently, multiple pathogenic mechanisms have been suggested including haploinsufficiency, truncated titin polypeptides as well as post-translational modifications of titin [99,100]. The role of missense variants is poorly understood, but at least for DCM their relevance as causative remains questionable; they may have a modifying effect [101]. Rarely, other cardiac phenotypes such as HCM, RCM and ACM have been suggested to be associated with TTN variants. In particular a de novo missense mutation, p.Y7621C, located in the A/I junction of titin has been shown to segregate in a family with five affected members aged 12–35 years with typical features of a restrictive physiology suggesting that other missense mutations may also relevant for RCM in particular if they appear de novo [34].

3.2. Mutations in Genes Encoding Non-Saromere Proteins

Although the majority of RCM-associated mutations has been found in genes encoding for different sarcomere proteins (Figure 2A), mutations in non-sarcomeric genes are also relevant. Several different mutations have been reported, for example in the DES and FLNC genes.

3.2.1. Desmin (DES)

The DES gene encodes the cytoplasmic muscle specific intermediate filament protein desmin. Intermediate filaments connect different cell organelles such as the cardiac desmosomes, costameres, Z-discs, mitochondria and the cell nuclei [102,103]. Cardiac desmosomes are cell–cell junctions localized at the intercalated disc mediating the cell–cell adhesion of the cardiomyocytes [104]. Desmin filaments are coupled to the desmosomes via the cytolinker protein, desmoplakin [105]. Costameres are multi-protein complexes localized at the sarcolemma and connect the extracellular matrix with the myofibrils [106]. The intermediate filaments are connected via different cytolinker proteins, e.g., plectin with the Z-bands and the costameres [107]. Due to its central role in the cardiac intermediate filament system and its connections with several multi-protein complexes or cell organelles, desmin is highly relevant for the structural integrity of the cardiomyocytes. DES-deficient mice developed severe cardiomyopathy in combination with skeletal myopathy characterized by fragile myofibrils, severe cardiac fibrosis, cardiomyocyte necrosis and abnormal calcium deposits [108,109]. DES mutations in humans are associated with different skeletal and cardiac myopathies [110,111,112,113,114]. In 2006, Hager and colleagues described for the first time a patient with RCM carrying the mutation DES-p.E245D. Later, it was recognized that this mutation causes a splicing defect leading to an in-frame skipping of exon-3 causing a deletion of 32 amino acids within the rod domain [115,116]. Several other pathogenic RCM-associated DES mutations have been reported [14,117,118,119,120,121,122] (Figure 5 and Table 4).
Most of the DES mutations are missense or small in-frame deletion mutations leading to a detrimental effect on the filament assembly process [123,124]. The desmin monomer consists of a central α-helical rod domain flanked by non-helical head and tail domains [125]. Two desmin monomers form coiled–coil dimers driven by the annealing of a hydrophobic seam [126]. These dimers form anti parallel tetramers [127]. Eight tetramers anneal into unit-length filaments (ULFs) which have a size of about 60 nm [128]. ULFs are the essential building blocks of intermediate filaments and hybridize longitudinally into regular intermediate filaments [125,129]. As intermediate filaments do not have a polar orientation, they can fuse end-to-end [130,131,132]. DES mutations can disturb the filament assembly at different steps [123,124].
Table 4. Overview about known RCM-associated DES mutations.
Table 4. Overview about known RCM-associated DES mutations.
MutationAge of Onset and Clinical FeaturesFamily HistoryMAF 1CommentsReferences
c.735+1G>Aadult, SMde novo-induces a splice defect, skipping of exon-3[133]
c.735+1G>Tadults, SMtwo patients-induces a splice defect, skipping of exon-3[119]
p.R16Cadult, AVB, HTxone patient0.000006570homozygous[134]
p.Y122Hadult, AVBone patient-homozygous[14]
c.735G>C
(p.E245D)
adults, AFseveral family members, only index patient was genotyped -induces a splice defect, skipping of exon-3[116]
p.I367Fadults, AVB, SMseveral family members-index patient diagnosed with HCM [135][15,135]
p.L392Padult, AVB, SMone patient- [135]
p.R406Wadults, AVBthree affected members-a different index patient presented ACM in combination with SM [112][117,134]
p.E413Kadults, AVB, AF, SCDfour affected members- [136,137]
p.R415Qadult, AFseveral family members-different phenotypes,
unclear if a splice defect is caused (last bp of exon-6)
[15]
p.P419Sadults, AVB, SMtwo patients- [135]
p.P433Tadult, AVB, SMone patient- [120]
p.T453Iadult, AVBde novo- [134]
p.R454Wadults, AVB, SMtwo patients- [112]
1 MAF = Minor allele frequency according to Genome Aggregation Database, https://gnomad.broadinstitute.org/ (accessed on 13 March 2022). ACM = arrhythmogenic cardiomyopathy, AF = atrial fibrillation, AVB = atrioventricular block, HCM = hypertrophic cardiomyopathy, HTx = heart transplantation, SCD = sudden cardiac death, SM = skeletal myopathy.

3.2.2. Myopalladin (MYPN)

Myopalladin belongs beside myotillin (MYOT) and palladin (PALLD) to the actin-binding and immunoglobulin-containing proteins within the Z-disc [138,139]. It contains five immunoglobulin (Ig) domains and a proline-rich motif [138]. In 2012, Purevjav et al. described a MYPN nonsense mutation (p.Q529X) in two affected siblings with RCM [33]. Beside RCM, MYPN mutations are also found in patients with DCM [140], HCM [141] and nemaline myopathy (MIM, #617336) [142].

3.2.3. α-Actinin-2 (ACTN2)

The ACTN2 gene was mapped to chromosome 1q43 [143] and consists of 21 exons [144]. α-Actinin-2 is the main structural component of the Z-discs in striated muscles [145] and belongs to the spectrin protein family [146]. The typical structural element of this protein family are the spectrin-like repeats [147], which are formed by three α-helices forming a left-handed supercoil [148]. α-Actinin-2 forms anti parallel dimers and consists of an N-terminal actin binding domain, a central ROD domain and a calmodulin-like domain (CAMD) (Figure 6) [149].
In 2016, Kostareva et al. screened a cohort of 24 unrelated RCM patients using a broad cardiomyopathy next generation sequencing (NGS) panel and identified, among others, the likely pathogenic mutation ACTN2-p.N175Y (Table 1) [39]. Besides RCM, pathogenic mutations in ACTN2 are associated with DCM [150], HCM [151], LVNC [152] or ACM [153] indicating a broad spectrum of cardiac phenotypes associated with those mutations (Figure 3). In addition, ACTN2 mutations can also cause skeletal myopathies [154].
Figure 6. Structural overview of the anti parallel α-actinin-2 dimer (https://www.rcsb.org/structure/4D1E) (accessed on 13 March 2022) [149]. The N-terminal. Actin-binding domains are shown in red. Four spectrin-like repeats build the central cylindrical rod domain (green). A C-terminal calmodulin-like domain is built by two EF hand motifs (purple and blue). The position of the RCM-associated mutation ACTN2-p.N157Y within the actin-binding domain is shown in cyan.
Figure 6. Structural overview of the anti parallel α-actinin-2 dimer (https://www.rcsb.org/structure/4D1E) (accessed on 13 March 2022) [149]. The N-terminal. Actin-binding domains are shown in red. Four spectrin-like repeats build the central cylindrical rod domain (green). A C-terminal calmodulin-like domain is built by two EF hand motifs (purple and blue). The position of the RCM-associated mutation ACTN2-p.N157Y within the actin-binding domain is shown in cyan.
Jcm 11 02094 g006

3.2.4. Filamin-C (FLNC)

Originally, mutations in FLNC were identified in patients with MFM (MIM, #609524) [155] or distal myopathy (MIM, #614065) [156]. The FLNC gene consists of 48 exons and is mapped on human chromosome 7q32 [157]. It encodes filamin-C, which is a cytolinker protein. Filamin-C contains an N-terminal actin-binding domain and 24 immunoglobulin-(Ig) domains, which are separated by two hinge regions (Figure 7) [158]. The dimerization of filamin-C is mediated by a protein–protein interaction of its 24th Ig-domains [159]. In cardiomyocytes, filamin-C is localized at the intercalated discs, the sarcolemma and the Z-discs [158,160]. Several binding partners including titin [161,162], integrin β1A and myotilin [163] as well as actin and sarcoglycans [164] have been reported. For a detailed overview see [158].
Valdés-Mas et al. identified in 2014, by whole-exome sequencing several FLNC mutations in patients with HCM [165]. Of note, FLNC mutations can likewise cause DCM [166], ACM [167] or non-compaction cardiomyopathy [168]. RCM-associated FLNC mutations were described in two families for the first time in 2016 [37]. Since then, several other FLNC missense mutations have been identified in RCM patients (Table 5 and Figure 7). Whereas DCM-associated FLNC truncation mutations are presumably leading to haploinsufficiency [169], an abnormal aggregation leading in consequence to sarcomeric disarray has been demonstrated for some missense mutations [37,165,170]. Several (zebra)fish and mouse models for FLNC have been generated revealing different muscle and heart defects [166,171,172,173,174,175,176,177,178]. Of note, even in Drosophila melanogaster loss of the filamin ortholog ‘Cheerio’ causes Z-disc and sarcomere defects [162]. Recently, two iPSC lines from donors with RCM carrying FLNC missense mutations have been generated, but their characterization is ongoing [179,180]. Tucker et al. inserted the mutation FLNC-p.V2297M using genome editing by ‘Clustered Regularly Interspaced Short Palindromic Repeats’ (CRISPR)-Cas9 into a human embryonic stem cell line (hESC). The fractional shortening was decreased in hESC-derived cardiomyocytes [181].

3.2.5. Lamin A/C (LMNA)

Lamin A/C belongs to the intermediate filament protein family (type V) [125] and forms the nuclear lamina [185]. The nuclear lamina is a molecular meshwork, which is important for the structural integrity of the nuclei and regulates the chromatin organization [186].
Recently, Paller et al. found a 1 bp deletion in exon 5 of the LMNA gene (c.835delG, p.E279RfsX201) in a RCM patient who developed additionally skeletal muscle weakness and atrial fibrillation [41]. Histology analysis revealed hypertrophy and cardiac fibrosis in the explanted myocardial tissue [41]. Beside RCM, LMNA mutations cause DCM [187], ACM [188], LVNC [189], Emery–Dreifuss muscular dystrophy (MIM, #181350) [190], familial lipodystrophy (MIM #151660) [191] and Hutchinson–Gilford progeria syndrome (HGPS, MIM #176670) [192]. The nuclear envelope and the connected nuclear lamina of cardiomyocytes are sensitive structures where mutations affect several other proteins, e.g., TMEM43 may cause different cardiomyopathies [193].

3.2.6. Transmembrane Protein 87B (TMEM87B)

TMEM87B encodes a multi-pass transmembrane protein, which is involved in endosome to Golgi apparatus retrograde transport [194].
Yu et al. described the hemizygous missense mutation TMEM87B-p.N456D in combination with a 1.7 Mb microdeletion on the second allele in a patient who developed RCM in combination with an atrial septal defect, craniofacial abnormalities, dysmorphic features, microcephaly and skeletal dysplasia [38]. Using anti sense morpholino injections, it has been shown by Russel et al. that TMEM87B knockdown causes cardiac hypoplasia and cardiac defects in zebrafish embryos [195].

3.2.7. αB-Crystallin (CRYAB)

CRYAB (or HSPB5) encodes αB-crystallin, which belongs to the small heat shock protein (sHSP) family [196]. Several sHSPs are expressed in the human heart. Originally, αB-crystallin was discovered as a major component of the vertebrate eye lenses [197]. However, it is also highly expressed in the heart and in the skeletal muscle [198,199]. In 1998, Vicart et al. identified in a French family with MFM in combination with HCM and cataract the pathogenic missense mutation CRYAB-p.R120G (Figure 8). Of note, this mutation causes, comparable to DES mutation, an abnormal aggregation of desmin and αB-crystallin in skeletal and cardiac myocytes [19]. Interestingly, Sacconi et al. described the same triad of clinical symptoms in a family carrying a different CRYAB mutation [200]. CRYAB mutations can also cause isolated cataract without cardiac involvement [201] or vice versa isolated DCM without cataract [202]. Recently, the CRYAB mutation p.D109G has been described in a small German family with RCM in combination with SM [40]. Interestingly, R120 and D109 form two ion bridges stabilizing the dimerization of αB-crystallin (Figure 8). The αB-crystallin dimers form large oligomers [203] which have an ATP-independent chaperone-like activity [204]. In addition, αB-crystallin binds also to different cytoskeletal and sarcomere proteins, e.g., titin [205].

3.2.8. Bcl2 Associated Athanogene 3 (BAG3)

The BAG3 gene consists of four exons and encodes Bcl2 associated athanogene 3 [208]. BAG3 is a co-chaperone binding to the ATPase domain of heat shock protein Hsc70/Hsp70 and regulating its chaperone function [209]. BAG3 is structurally organized in an N-terminal tryptophan-tryptophan (WW) domain, two IPV domains, two 14-3-3 binding motifs, a proline-rich region and a C-terminal BAG domain [210,211]. The protein–protein interaction of BAG3 with Hsc70/Hsp70 is mediated by its BAG domain [212]. BAG3 acts as an ATP exchange factor stabilizing the ATPase domain of Hsc70/Hsp70 without bound ATP [213]. Since the multi-domain organization of BAG3, numerous other binding partners have been described. For example, BAG3 binds to several members of the sHSP family including αB-Crystallin [214,215,216,217,218]. Briefly summarized, BAG3 has a central and important role in protein quality control and chaperone-assisted selective autophagy [219].
Several pathogenic mutations in BAG3 have been described in patients with DCM [220] or with MFM [221]. In addition, BAG3 mutations are found in patients with RCM in combination with MFM [42]. Recently, Kimura et al. generated a transgenic mouse model with an overexpression of BAG3-p.P209L conjugated with green fluorescent protein. These mice develop RCM and severe cardiac fibrosis. At the cellular level, disorganization of the Z-disc and abnormal protein aggregation were present [222]. In contrast, the knock-in mouse model carrying the equivalent murine mutation Bag3-p.P215L does not develop a cardiac phenotype [223].

3.2.9. Discoidin Cub and Lccl Domain Containing Protein-2 (DCBLD2)

Recently, Alhamoudi et al. described the homozygous nonsense mutation DCBLD2-p.W27X in a 5-year-old Arabic patient with severe RCM, tachycardia, atrial fibrillation, dysmorphic features and developmental delay. Functional analyses using primary dermal fibroblast from the mutation carrier indicated reduced cell proliferation and altered amounts of calcium and reactive oxygen species in comparison to normal fibroblasts [43]. DCBLD2 encodes a ubiquitously expressed type-I transmembrane protein [224,225]. It is involved in vascular smooth muscle cell proliferation [226], vascular endothelial growth factor (VEGF) signaling [227] and epithelial–mesenchymal transition [228]. However, the exact molecular functions of DCBLD2 contributing to RCM and other cardiomyopathies are currently unknown and deserve increased research attention in the future.

4. Summary and Outlook

Currently, mutations in over 19 different disease-causing genes have been discovered in patients with primary RCM. However, the genetic landscape of RCM is overlapping with the genetic background of other cardiomyopathies. Genes encoding for sarcomere proteins such as cardiac troponin-I are the major RCM genes. However, more recently, the prevalence of mutations in specific non-sarcomeric genes such as DES or FLNC has increased; broad NGS gene panels or whole exome sequencing should be considered if a genetic etiology is suspected. This might be also beneficial, since the genetic landscape of RCM remains incomplete. Therefore, multi-center studies enrolling larger patient cohorts are needed to provide a robust overview about the genetic etiology of RCM. In addition, these studies might reveal the age of onset associated with specific genotypes.
As no sufficient treatment for RCM is currently available, there is a highly unmet medical need for the development of more precise genetic or molecular therapies. However, there is hope on the horizon with novel therapies targeting the sarcomere. In particular, for the obstructive form of HCM the allosteric inhibitor of the cardiac specific myosin adenosine triphosphatase (MYK-461) has shown symptomatic improvement in a phase 3 trial and may also be applicable for patients with RCM and sarcomeric mutations leading to an excessive cross bridging with actin [229]. The opposite setting, small molecules, such as omecamtiv mecarbil and danicamtiv, increasing contractility may be effective in particular in patients with sarcomere mutations and DCM [230].
Another exciting strategy can be seen in genome editing using CRISPR-Cas9 [231] or RNA editing using Cas7-11 [232] in combination with adequate cardiomyocyte specific delivery vectors, e.g., adeno-associated viruses [233,234], will help to reach this goal in the future. Recently, CRISPR-Cas9 has been used for example for correcting DCM associated truncating TTN mutations [235] and deserves interest in the context of RCM in the future.

Author Contributions

Conceptualization, A.B. and B.G.; writing—original draft preparation, A.B. and B.G.; writing—review and editing, A.B. and B.G.; visualization, A.B. and B.G.; funding acquisition, A.B. and B.G. All authors have read and agreed to the published version of the manuscript.

Funding

We acknowledge support by the DFG Open Access Publication Funds of the Ruhr-University Bochum (funding number: 5678). Funding to B.G. was also provided by the German Research Foundation (project number: 453989101).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. McKenna, W.J.; Maron, B.J.; Thiene, G. Classification, epidemiology, and global burden of cardiomyopathies. Circ. Res. 2017, 121, 722–730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. White, P.; Myers, M. The classification of cardiac diagnosis. JAMA 1921, 77, 1414–1415. [Google Scholar]
  3. Gerull, B.; Klaassen, S.; Brodehl, A. The genetic landscape of cardiomyopathies. In Genetic Causes of Cardiac Disease; Erdmann, J., Moretti, A., Eds.; Springer: Cham, Switzerland, 2019; pp. 45–91. [Google Scholar]
  4. Butzner, M.; Leslie, D.L.; Cuffee, Y.; Hollenbeak, C.S.; Sciamanna, C.; Abraham, T. Stable rates of obstructive hypertrophic cardiomyopathy in a contemporary era. Front. Cardiovasc. Med. 2021, 8, 765876. [Google Scholar] [CrossRef]
  5. Elliott, P.; Andersson, B.; Arbustini, E.; Bilinska, Z.; Cecchi, F.; Charron, P.; Dubourg, O.; Kuhl, U.; Maisch, B.; McKenna, W.J.; et al. Classification of the cardiomyopathies: A position statement from the european society of cardiology working group on myocardial and pericardial diseases. Eur. Heart J. 2008, 29, 270–276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Cuddy, S.A.M.; Falk, R.H. Amyloidosis as a systemic disease in context. Can. J. Cardiol. 2020, 36, 396–407. [Google Scholar] [CrossRef] [PubMed]
  7. Ruberg, F.L.; Grogan, M.; Hanna, M.; Kelly, J.W.; Maurer, M.S. Transthyretin amyloid cardiomyopathy: JACC state-of-the-art review. J. Am. Coll. Cardiol. 2019, 73, 2872–2891. [Google Scholar] [CrossRef] [PubMed]
  8. Paisey, R.B.; Steeds, R.; Barrett, T.; Williams, D.; Geberhiwot, T.; Gunay-Aygun, M. Alstrom syndrome. In GeneReviews((R)); Adam, M.P., Ardinger, H.H., Pagon, R.A., Wallace, S.E., Bean, L.J.H., Gripp, K.W., Mirzaa, G.M., Amemiya, A., Eds.; University of Washington: Seattle, DC, USA, 1993. [Google Scholar]
  9. Starr, L.J.; Lindor, N.M.; Lin, A.E. Myhre syndrome. In GeneReviews((R)); Adam, M.P., Ardinger, H.H., Pagon, R.A., Wallace, S.E., Bean, L.J.H., Gripp, K.W., Mirzaa, G.M., Amemiya, A., Eds.; University of Washington: Seattle, DC, USA, 1993. [Google Scholar]
  10. Seferovic, P.M.; Polovina, M.; Bauersachs, J.; Arad, M.; Gal, T.B.; Lund, L.H.; Felix, S.B.; Arbustini, E.; Caforio, A.L.P.; Farmakis, D.; et al. Heart failure in cardiomyopathies: A position paper from the heart failure association of the european society of cardiology. Eur. J. Heart Fail. 2019, 21, 553–576. [Google Scholar] [CrossRef] [Green Version]
  11. Kubo, T.; Gimeno, J.R.; Bahl, A.; Steffensen, U.; Steffensen, M.; Osman, E.; Thaman, R.; Mogensen, J.; Elliott, P.M.; Doi, Y.; et al. Prevalence, clinical significance, and genetic basis of hypertrophic cardiomyopathy with restrictive phenotype. J. Am. Coll. Cardiol. 2007, 49, 2419–2426. [Google Scholar] [CrossRef] [Green Version]
  12. DePasquale, E.C.; Nasir, K.; Jacoby, D.L. Outcomes of adults with restrictive cardiomyopathy after heart transplantation. J. Heart Lung Transplant. 2012, 31, 1269–1275. [Google Scholar] [CrossRef]
  13. Muchtar, E.; Blauwet, L.A.; Gertz, M.A. Restrictive cardiomyopathy: Genetics, pathogenesis, clinical manifestations, diagnosis, and therapy. Circ. Res. 2017, 121, 819–837. [Google Scholar] [CrossRef]
  14. Brodehl, A.; Pour Hakimi, S.A.; Stanasiuk, C.; Ratnavadivel, S.; Hendig, D.; Gaertner, A.; Gerull, B.; Gummert, J.; Paluszkiewicz, L.; Milting, H. Restrictive cardiomyopathy is caused by a novel homozygous desmin (DES) mutation p.Y122H leading to a severe filament assembly defect. Genes 2019, 10, 918. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Ripoll-Vera, T.; Zorio, E.; Gamez, J.M.; Molina, P.; Govea, N.; Cremer, D. Phenotypic patterns of cardiomyopathy caused by mutations in the desmin gene. A clinical and genetic study in two inherited heart disease units. Rev. Esp. Cardiol. 2015, 68, 1027–1029. [Google Scholar] [CrossRef] [PubMed]
  16. Dorsch, L.M.; Kuster, D.W.D.; Jongbloed, J.D.H.; Boven, L.G.; van Spaendonck-Zwarts, K.Y.; Suurmeijer, A.J.H.; Vink, A.; du Marchie Sarvaas, G.J.; van den Berg, M.P.; van der Velden, J.; et al. The effect of tropomyosin variants on cardiomyocyte function and structure that underlie different clinical cardiomyopathy phenotypes. Int. J. Cardiol. 2021, 323, 251–258. [Google Scholar] [CrossRef] [PubMed]
  17. Fichna, J.P.; Maruszak, A.; Zekanowski, C. Myofibrillar myopathy in the genomic context. J. Appl. Genet. 2018, 59, 431–439. [Google Scholar] [CrossRef] [Green Version]
  18. Munoz-Marmol, A.M.; Strasser, G.; Isamat, M.; Coulombe, P.A.; Yang, Y.; Roca, X.; Vela, E.; Mate, J.L.; Coll, J.; Fernandez-Figueras, M.T.; et al. A dysfunctional desmin mutation in a patient with severe generalized myopathy. Proc. Natl. Acad. Sci. USA 1998, 95, 11312–11317. [Google Scholar] [CrossRef] [Green Version]
  19. Vicart, P.; Caron, A.; Guicheney, P.; Li, Z.; Prevost, M.C.; Faure, A.; Chateau, D.; Chapon, F.; Tome, F.; Dupret, J.M.; et al. A missense mutation in the αB-crystallin chaperone gene causes a desmin-related myopathy. Nat. Genet. 1998, 20, 92–95. [Google Scholar] [CrossRef]
  20. Kley, R.A.; Hellenbroich, Y.; van der Ven, P.F.; Furst, D.O.; Huebner, A.; Bruchertseifer, V.; Peters, S.A.; Heyer, C.M.; Kirschner, J.; Schroder, R.; et al. Clinical and morphological phenotype of the filamin myopathy: A study of 31 German patients. Brain 2007, 130 Pt 12, 3250–3264. [Google Scholar] [CrossRef]
  21. Dhawan, P.S.; Liewluck, T.; Knapik, J.; Milone, M. Myofibrillar myopathy due to dominant LMNA mutations: A report of 2 cases. Muscle Nerve 2018, 57, E124–E126. [Google Scholar] [CrossRef]
  22. Odgerel, Z.; Sarkozy, A.; Lee, H.S.; McKenna, C.; Rankin, J.; Straub, V.; Lochmuller, H.; Paola, F.; D’Amico, A.; Bertini, E.; et al. Inheritance patterns and phenotypic features of myofibrillar myopathy associated with a BAG3 mutation. Neuromuscul. Disord. 2010, 20, 438–442. [Google Scholar] [CrossRef] [Green Version]
  23. Pfeffer, G.; Barresi, R.; Wilson, I.J.; Hardy, S.A.; Griffin, H.; Hudson, J.; Elliott, H.R.; Ramesh, A.V.; Radunovic, A.; Winer, J.B.; et al. Titin founder mutation is a common cause of myofibrillar myopathy with early respiratory failure. J. Neurol. Neurosurg. Psychiatry 2014, 85, 331–338. [Google Scholar] [CrossRef]
  24. Izumi, R.; Niihori, T.; Aoki, Y.; Suzuki, N.; Kato, M.; Warita, H.; Takahashi, T.; Tateyama, M.; Nagashima, T.; Funayama, R.; et al. Exome sequencing identifies a novel TTN mutation in a family with hereditary myopathy with early respiratory failure. J. Hum. Genet. 2013, 58, 259–266. [Google Scholar] [CrossRef] [PubMed]
  25. Weterman, M.A.; Barth, P.G.; van Spaendonck-Zwarts, K.Y.; Aronica, E.; Poll-The, B.T.; Brouwer, O.F.; van Tintelen, J.P.; Qahar, Z.; Bradley, E.J.; de Wissel, M.; et al. Recessive MYL2 mutations cause infantile type I muscle fibre disease and cardiomyopathy. Brain 2013, 136 Pt 1, 282–293. [Google Scholar] [CrossRef] [Green Version]
  26. Cimiotti, D.; Budde, H.; Hassoun, R.; Jaquet, K. Genetic restrictive cardiomyopathy: Causes and consequences—An integrative approach. Int. J. Mol. Sci. 2021, 22, 558. [Google Scholar] [CrossRef]
  27. Mogensen, J.; Kubo, T.; Duque, M.; Uribe, W.; Shaw, A.; Murphy, R.; Gimeno, J.R.; Elliott, P.; McKenna, W.J. Idiopathic restrictive cardiomyopathy is part of the clinical expression of cardiac troponin I mutations. J. Clin. Investig. 2003, 111, 209–216. [Google Scholar] [CrossRef] [Green Version]
  28. Peddy, S.B.; Vricella, L.A.; Crosson, J.E.; Oswald, G.L.; Cohn, R.D.; Cameron, D.E.; Valle, D.; Loeys, B.L. Infantile restrictive cardiomyopathy resulting from a mutation in the cardiac troponin T gene. Pediatrics 2006, 117, 1830–1833. [Google Scholar] [CrossRef] [PubMed]
  29. Hager, S.; Mahrholdt, H.; Goldfarb, L.G.; Goebel, H.H.; Sechtem, U. Images in cardiovascular medicine. Giant right atrium in the setting of desmin-related restrictive cardiomyopathy. Circulation 2006, 113, e53–e55. [Google Scholar] [CrossRef] [PubMed]
  30. Kaski, J.P.; Syrris, P.; Burch, M.; Tome-Esteban, M.T.; Fenton, M.; Christiansen, M.; Andersen, P.S.; Sebire, N.; Ashworth, M.; Deanfield, J.E.; et al. Idiopathic restrictive cardiomyopathy in children is caused by mutations in cardiac sarcomere protein genes. Heart 2008, 94, 1478–1484. [Google Scholar] [CrossRef]
  31. Karam, S.; Raboisson, M.J.; Ducreux, C.; Chalabreysse, L.; Millat, G.; Bozio, A.; Bouvagnet, P. A de novo mutation of the beta cardiac myosin heavy chain gene in an infantile restrictive cardiomyopathy. Congenit. Heart Dis. 2008, 3, 138–143. [Google Scholar] [CrossRef]
  32. Caleshu, C.; Sakhuja, R.; Nussbaum, R.L.; Schiller, N.B.; Ursell, P.C.; Eng, C.; De Marco, T.; McGlothlin, D.; Burchard, E.G.; Rame, J.E. Furthering the link between the sarcomere and primary cardiomyopathies: Restrictive cardiomyopathy associated with multiple mutations in genes previously associated with hypertrophic or dilated cardiomyopathy. Am. J. Med. Genet. Part A 2011, 155, 2229–2235. [Google Scholar] [CrossRef] [Green Version]
  33. Purevjav, E.; Arimura, T.; Augustin, S.; Huby, A.C.; Takagi, K.; Nunoda, S.; Kearney, D.L.; Taylor, M.D.; Terasaki, F.; Bos, J.M.; et al. Molecular basis for clinical heterogeneity in inherited cardiomyopathies due to myopalladin mutations. Hum. Mol. Genet. 2012, 21, 2039–2053. [Google Scholar] [CrossRef] [Green Version]
  34. Peled, Y.; Gramlich, M.; Yoskovitz, G.; Feinberg, M.S.; Afek, A.; Polak-Charcon, S.; Pras, E.; Sela, B.A.; Konen, E.; Weissbrod, O.; et al. Titin mutation in familial restrictive cardiomyopathy. Int. J. Cardiol. 2014, 171, 24–30. [Google Scholar] [CrossRef] [PubMed]
  35. Wu, W.; Lu, C.X.; Wang, Y.N.; Liu, F.; Chen, W.; Liu, Y.T.; Han, Y.C.; Cao, J.; Zhang, S.Y.; Zhang, X. Novel phenotype-genotype correlations of restrictive cardiomyopathy with myosin-binding protein C (MYBPC3) gene mutations tested by next-generation sequencing. J. Am. Heart Assoc. 2015, 4, e001879. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Ploski, R.; Rydzanicz, M.; Ksiazczyk, T.M.; Franaszczyk, M.; Pollak, A.; Kosinska, J.; Michalak, E.; Stawinski, P.; Ziolkowska, L.; Bilinska, Z.T.; et al. Evidence for troponin C (TNNC1) as a gene for autosomal recessive restrictive cardiomyopathy with fatal outcome in infancy. Am. J. Med. Genet. Part A 2016, 170, 3241–3248. [Google Scholar] [CrossRef] [PubMed]
  37. Brodehl, A.; Ferrier, R.A.; Hamilton, S.J.; Greenway, S.C.; Brundler, M.A.; Yu, W.; Gibson, W.T.; McKinnon, M.L.; McGillivray, B.; Alvarez, N.; et al. Mutations in FLNC are associated with familial restrictive cardiomyopathy. Hum. Mutat. 2016, 37, 269–279. [Google Scholar] [CrossRef]
  38. Yu, H.C.; Coughlin, C.R.; Geiger, E.A.; Salvador, B.J.; Elias, E.R.; Cavanaugh, J.L.; Chatfield, K.C.; Miyamoto, S.D.; Shaikh, T.H. Discovery of a potentially deleterious variant in TMEM87B in a patient with a hemizygous 2q13 microdeletion suggests a recessive condition characterized by congenital heart disease and restrictive cardiomyopathy. Mol. Case Stud. 2016, 2, a000844. [Google Scholar] [CrossRef] [Green Version]
  39. Kostareva, A.; Kiselev, A.; Gudkova, A.; Frishman, G.; Ruepp, A.; Frishman, D.; Smolina, N.; Tarnovskaya, S.; Nilsson, D.; Zlotina, A.; et al. Genetic spectrum of idiopathic restrictive cardiomyopathy uncovered by next-generation sequencing. PLoS ONE 2016, 11, e0163362. [Google Scholar] [CrossRef] [Green Version]
  40. Brodehl, A.; Gaertner-Rommel, A.; Klauke, B.; Grewe, S.A.; Schirmer, I.; Peterschroder, A.; Faber, L.; Vorgerd, M.; Gummert, J.; Anselmetti, D.; et al. The novel αB-crystallin (CRYAB) mutation p.D109G causes restrictive cardiomyopathy. Hum. Mutat. 2017, 38, 947–952. [Google Scholar] [CrossRef]
  41. Paller, M.S.; Martin, C.M.; Pierpont, M.E. Restrictive cardiomyopathy: An unusual phenotype of a lamin A variant. ESC Heart Fail. 2018, 5, 724–726. [Google Scholar] [CrossRef]
  42. Schanzer, A.; Rupp, S.; Graf, S.; Zengeler, D.; Jux, C.; Akinturk, H.; Gulatz, L.; Mazhari, N.; Acker, T.; Van Coster, R.; et al. Dysregulated autophagy in restrictive cardiomyopathy due to Pro209Leu mutation in BAG3. Mol. Genet. Metab. 2018, 123, 388–399. [Google Scholar] [CrossRef]
  43. Alhamoudi, K.M.; Barhoumi, T.; Al-Eidi, H.; Asiri, A.; Nashabat, M.; Alaamery, M.; Alharbi, M.; Alhaidan, Y.; Tabarki, B.; Umair, M.; et al. A homozygous nonsense mutation in DCBLD2 is a candidate cause of developmental delay, dysmorphic features and restrictive cardiomyopathy. Sci. Rep. 2021, 11, 12861. [Google Scholar] [CrossRef]
  44. Gordon, A.M.; Homsher, E.; Regnier, M. Regulation of contraction in striated muscle. Physiol. Rev. 2000, 80, 853–924. [Google Scholar] [CrossRef] [PubMed]
  45. Thierfelder, L.; Watkins, H.; MacRae, C.; Lamas, R.; McKenna, W.; Vosberg, H.P.; Seidman, J.G.; Seidman, C.E. α-tropomyosin and cardiac troponin T mutations cause familial hypertrophic cardiomyopathy: A disease of the sarcomere. Cell 1994, 77, 701–712. [Google Scholar] [CrossRef]
  46. Gomes, A.V.; Liang, J.; Potter, J.D. Mutations in human cardiac troponin I that are associated with restrictive cardiomyopathy affect basal ATPase activity and the calcium sensitivity of force development. J. Biol. Chem. 2005, 280, 30909–30915. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Cimiotti, D.; Fujita-Becker, S.; Mohner, D.; Smolina, N.; Budde, H.; Wies, A.; Morgenstern, L.; Gudkova, A.; Sejersen, T.; Sjoberg, G.; et al. Infantile restrictive cardiomyopathy: cTnI-R170G/W impair the interplay of sarcomeric proteins and the integrity of thin filaments. PLoS ONE 2020, 15, e0229227. [Google Scholar] [CrossRef] [PubMed]
  48. Davis, J.; Wen, H.; Edwards, T.; Metzger, J.M. Allele and species dependent contractile defects by restrictive and hypertrophic cardiomyopathy-linked troponin I mutants. J. Mol. Cell. Cardiol. 2008, 44, 891–904. [Google Scholar] [CrossRef] [Green Version]
  49. Li, Y.; Charles, P.Y.; Nan, C.; Pinto, J.R.; Wang, Y.; Liang, J.; Wu, G.; Tian, J.; Feng, H.Z.; Potter, J.D.; et al. Correcting diastolic dysfunction by Ca2+ desensitizing troponin in a transgenic mouse model of restrictive cardiomyopathy. J. Mol. Cell. Cardiol. 2010, 49, 402–411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Li, Y.; Zhang, L.; Jean-Charles, P.Y.; Nan, C.; Chen, G.; Tian, J.; Jin, J.P.; Gelb, I.J.; Huang, X. Dose-dependent diastolic dysfunction and early death in a mouse model with cardiac troponin mutations. J. Mol. Cell. Cardiol. 2013, 62, 227–236. [Google Scholar] [CrossRef] [Green Version]
  51. Kawai, M.; Johnston, J.R.; Karam, T.; Wang, L.; Singh, R.K.; Pinto, J.R. Myosin rod hypophosphorylation and CB kinetics in papillary muscles from a TnC-A8V KI mouse model. Biophys. J. 2017, 112, 1726–1736. [Google Scholar] [CrossRef] [Green Version]
  52. Redwood, C.; Robinson, P. Alpha-tropomyosin mutations in inherited cardiomyopathies. J. Muscle Res. Cell Motil. 2013, 34, 285–294. [Google Scholar] [CrossRef]
  53. Hassoun, R.; Budde, H.; Mannherz, H.G.; Lódi, M.; Fujita-Becker, S.; Laser, K.T.; Gärtner, A.; Klingel, K.; Möhner, D.; Stehle, R.; et al. De novo missense mutations in TNNC1 and TNNI3 causing severe infantile cardiomyopathy affect myofilament structure and function and are modulated by troponin targeting agents. Int. J. Mol. Sci. 2021, 22, 9625. [Google Scholar] [CrossRef]
  54. Mouton, J.M.; Pellizzon, A.S.; Goosen, A.; Kinnear, C.J.; Herbst, P.G.; Brink, P.A.; Moolman-Smook, J.C. Diagnostic disparity and identification of two TNNI3 gene mutations, one novel and one arising de novo, in South African patients with restrictive cardiomyopathy and focal ventricular hypertrophy. Cardiovasc. J. Afr. 2015, 26, 63–69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Van den Wijngaard, A.; Volders, P.; Van Tintelen, J.P.; Jongbloed, J.D.; van den Berg, M.P.; Lekanne Deprez, R.H.; Mannens, M.M.; Hofmann, N.; Slegtenhorst, M.; Dooijes, D.; et al. Recurrent and founder mutations in the Netherlands: Cardiac Troponin I (TNNI3) gene mutations as a cause of severe forms of hypertrophic and restrictive cardiomyopathy. Neth. Heart J. 2011, 19, 344–351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Ruan, Y.P.; Lu, C.X.; Zhao, X.Y.; Liang, R.J.; Lian, H.; Routledge, M.; Wu, W.; Zhang, X.; Fan, Z.J. Restrictive cardiomyopathy resulting from a troponin I type 3 mutation in a Chinese family. Chin. Med. Sci. J. 2016, 31, 1–7. [Google Scholar] [CrossRef]
  57. Kostareva, A.; Gudkova, A.; Sjöberg, G.; Mörner, S.; Semernin, E.; Krutikov, A.; Shlyakhto, E.; Sejersen, T. Deletion in TNNI3 gene is associated with restrictive cardiomyopathy. Int. J. Cardiol. 2009, 131, 410–412. [Google Scholar] [CrossRef]
  58. Mogensen, J.; Hey, T.; Lambrecht, S. A Systematic Review of Phenotypic Features Associated With Cardiac Troponin I Mutations in Hereditary Cardiomyopathies. Can. J. Cardiol. 2015, 31, 1377–1385. [Google Scholar] [CrossRef]
  59. Ding, W.H.; Han, L.; Xiao, Y.Y.; Mo, Y.; Yang, J.; Wang, X.F.; Jin, M. Role of Whole-exome sequencing in phenotype classification and clinical treatment of pediatric restrictive cardiomyopathy. Chin. Med. J. 2017, 130, 2823–2828. [Google Scholar] [CrossRef]
  60. Rai, T.S.; Ahmad, S.; Ahluwalia, T.S.; Ahuja, M.; Bahl, A.; Saikia, U.N.; Singh, B.; Talwar, K.K.; Khullar, M. Genetic and clinical profile of Indian patients of idiopathic restrictive cardiomyopathy with and without hypertrophy. Mol. Cell. Biochem. 2009, 331, 187–192. [Google Scholar] [CrossRef]
  61. Gerhardt, T.; Monserrat, L.; Landmesser, U.; Poller, W. A novel Troponin I mutation associated with severe restrictive cardiomyopathy-a case report of a 27-year-old woman with fatigue. Eur. Heart J. Case Rep. 2022, 6, ytac053. [Google Scholar] [CrossRef]
  62. Shah, S.; Yogasundaram, H.; Basu, R.; Wang, F.; Paterson, D.I.; Alastalo, T.P.; Oudit, G.Y. Novel dominant-negative mutation in cardiac troponin I causes severe restrictive cardiomyopathy. Circ. Heart Fail. 2017, 10, e003820. [Google Scholar] [CrossRef]
  63. Pantou, M.P.; Gourzi, P.; Gkouziouta, A.; Armenis, I.; Kaklamanis, L.; Zygouri, C.; Constantoulakis, P.; Adamopoulos, S.; Degiannis, D. A case report of recessive restrictive cardiomyopathy caused by a novel mutation in cardiac troponin I (TNNI3). BMC Med. Genet. 2019, 20, 61. [Google Scholar] [CrossRef]
  64. Yang, S.W.; Hitz, M.P.; Andelfinger, G. Ventricular septal defect and restrictive cardiomyopathy in a paediatric TNNI3 mutation carrier. Cardiol. Young 2010, 20, 574–576. [Google Scholar] [CrossRef] [PubMed]
  65. Gambarin, F.I.; Tagliani, M.; Arbustini, E. Pure restrictive cardiomyopathy associated with cardiac troponin I gene mutation: Mismatch between the lack of hypertrophy and the presence of disarray. Heart 2008, 94, 1257. [Google Scholar] [CrossRef] [PubMed]
  66. Menon, S.C.; Michels, V.V.; Pellikka, P.A.; Ballew, J.D.; Karst, M.L.; Herron, K.J.; Nelson, S.M.; Rodeheffer, R.J.; Olson, T.M. Cardiac troponin T mutation in familial cardiomyopathy with variable remodeling and restrictive physiology. Clin. Genet. 2008, 74, 445–454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Ezekian, J.E.; Clippinger, S.R.; Garcia, J.M.; Yang, Q.; Denfield, S.; Jeewa, A.; Dreyer, W.J.; Zou, W.; Fan, Y.; Allen, H.D.; et al. Variant R94C in TNNT2-encoded troponin t predisposes to pediatric restrictive cardiomyopathy and sudden death through impaired thin filament relaxation resulting in myocardial diastolic dysfunction. J. Am. Heart Assoc. 2020, 9, e015111. [Google Scholar] [CrossRef] [PubMed]
  68. Yamada, Y.; Namba, K.; Fujii, T. Cardiac muscle thin filament structures reveal calcium regulatory mechanism. Nat. Commun. 2020, 11, 153. [Google Scholar] [CrossRef] [PubMed]
  69. Ueyama, H.; Inazawa, J.; Ariyama, T.; Nishino, H.; Ochiai, Y.; Ohkubo, I.; Miwa, T. Reexamination of chromosomal loci of human muscle actin genes by fluorescence in situ hybridization. Jpn. J. Hum. Genet. 1995, 40, 145–148. [Google Scholar] [CrossRef] [Green Version]
  70. Gunning, P.; Ponte, P.; Kedes, L.; Eddy, R.; Shows, T. Chromosomal location of the co-expressed human skeletal and cardiac actin genes. Proc. Natl. Acad. Sci. USA 1984, 81, 1813–1817. [Google Scholar] [CrossRef] [Green Version]
  71. Squire, J. Special issue: The actin-myosin interaction in muscle: Background and overview. Int. J. Mol. Sci. 2019, 20, 5715. [Google Scholar] [CrossRef] [Green Version]
  72. Olson, T.M.; Michels, V.V.; Thibodeau, S.N.; Tai, Y.S.; Keating, M.T. Actin mutations in dilated cardiomyopathy, a heritable form of heart failure. Science 1998, 280, 750–752. [Google Scholar] [CrossRef] [Green Version]
  73. Olson, T.M.; Doan, T.P.; Kishimoto, N.Y.; Whitby, F.G.; Ackerman, M.J.; Fananapazir, L. Inherited and de novo mutations in the cardiac actin gene cause hypertrophic cardiomyopathy. J. Mol. Cell. Cardiol. 2000, 32, 1687–1694. [Google Scholar] [CrossRef]
  74. Klaassen, S.; Probst, S.; Oechslin, E.; Gerull, B.; Krings, G.; Schuler, P.; Greutmann, M.; Hurlimann, D.; Yegitbasi, M.; Pons, L.; et al. Mutations in sarcomere protein genes in left ventricular noncompaction. Circulation 2008, 117, 2893–2901. [Google Scholar] [CrossRef] [PubMed]
  75. Greenway, S.C.; McLeod, R.; Hume, S.; Roslin, N.M.; Alvarez, N.; Giuffre, M.; Zhan, S.H.; Shen, Y.; Preuss, C.; Andelfinger, G.; et al. Exome sequencing identifies a novel variant in ACTC1 associated with familial atrial septal defect. Can. J. Cardiol. 2014, 30, 181–187. [Google Scholar] [CrossRef] [PubMed]
  76. Sheikh, F.; Lyon, R.C.; Chen, J. Getting the skinny on thick filament regulation in cardiac muscle biology and disease. Trends Cardiovasc. Med. 2014, 24, 133–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Warrick, H.M.; Spudich, J.A. Myosin structure and function in cell motility. Annu. Rev. Cell Biol. 1987, 3, 379–421. [Google Scholar] [CrossRef]
  78. Alamo, L.; Ware, J.S.; Pinto, A.; Gillilan, R.E.; Seidman, J.G.; Seidman, C.E.; Padron, R. Effects of myosin variants on interacting-heads motif explain distinct hypertrophic and dilated cardiomyopathy phenotypes. eLife 2017, 6, e24634. [Google Scholar] [CrossRef]
  79. Colegrave, M.; Peckham, M. Structural implications of beta-cardiac myosin heavy chain mutations in human disease. Anat. Rec. 2014, 297, 1670–1680. [Google Scholar] [CrossRef]
  80. Vale, R.D.; Milligan, R.A. The way things move: Looking under the hood of molecular motor proteins. Science 2000, 288, 88–95. [Google Scholar] [CrossRef] [Green Version]
  81. Trybus, K.M. Role of myosin light chains. J. Muscle Res. Cell. Motil. 1994, 15, 587–594. [Google Scholar] [CrossRef]
  82. Wolny, M.; Colegrave, M.; Colman, L.; White, E.; Knight, P.J.; Peckham, M. Cardiomyopathy mutations in the tail of beta-cardiac myosin modify the coiled-coil structure and affect integration into thick filaments in muscle sarcomeres in adult cardiomyocytes. J. Biol. Chem. 2013, 288, 31952–31962. [Google Scholar] [CrossRef] [Green Version]
  83. Geisterfer-Lowrance, A.A.; Kass, S.; Tanigawa, G.; Vosberg, H.P.; McKenna, W.; Seidman, C.E.; Seidman, J.G. A molecular basis for familial hypertrophic cardiomyopathy: A beta cardiac myosin heavy chain gene missense mutation. Cell 1990, 62, 999–1006. [Google Scholar] [CrossRef]
  84. Moller, D.V.; Andersen, P.S.; Hedley, P.; Ersboll, M.K.; Bundgaard, H.; Moolman-Smook, J.; Christiansen, M.; Kober, L. The role of sarcomere gene mutations in patients with idiopathic dilated cardiomyopathy. Eur. J. Hum. Genet. 2009, 17, 1241–1249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Ferradini, V.; Parca, L.; Martino, A.; Lanzillo, C.; Silvetti, E.; Calo, L.; Caselli, S.; Novelli, G.; Helmer-Citterich, M.; Sangiuolo, F.C.; et al. Variants in MHY7 gene cause arrhythmogenic cardiomyopathy. Genes 2021, 12, 793. [Google Scholar] [CrossRef] [PubMed]
  86. Olson, T.M.; Karst, M.L.; Whitby, F.G.; Driscoll, D.J. Myosin light chain mutation causes autosomal recessive cardiomyopathy with mid-cavitary hypertrophy and restrictive physiology. Circulation 2002, 105, 2337–2340. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Yuan, C.C.; Kazmierczak, K.; Liang, J.; Kanashiro-Takeuchi, R.; Irving, T.C.; Gomes, A.V.; Wang, Y.; Burghardt, T.P.; Szczesna-Cordary, D. Hypercontractile mutant of ventricular myosin essential light chain leads to disruption of sarcomeric structure and function and results in restrictive cardiomyopathy in mice. Cardiovasc. Res. 2017, 113, 1124–1136. [Google Scholar] [CrossRef] [Green Version]
  88. Poetter, K.; Jiang, H.; Hassanzadeh, S.; Master, S.R.; Chang, A.; Dalakas, M.C.; Rayment, I.; Sellers, J.R.; Fananapazir, L.; Epstein, N.D. Mutations in either the essential or regulatory light chains of myosin are associated with a rare myopathy in human heart and skeletal muscle. Nat. Genet. 1996, 13, 63–69. [Google Scholar] [CrossRef]
  89. Flavigny, J.; Richard, P.; Isnard, R.; Carrier, L.; Charron, P.; Bonne, G.; Forissier, J.F.; Desnos, M.; Dubourg, O.; Komajda, M.; et al. Identification of two novel mutations in the ventricular regulatory myosin light chain gene (MYL2) associated with familial and classical forms of hypertrophic cardiomyopathy. J. Mol. Med. 1998, 76, 208–214. [Google Scholar] [CrossRef]
  90. Osborn, D.P.S.; Emrahi, L.; Clayton, J.; Tabrizi, M.T.; Wan, A.Y.B.; Maroofian, R.; Yazdchi, M.; Garcia, M.L.E.; Galehdari, H.; Hesse, C.; et al. Autosomal recessive cardiomyopathy and sudden cardiac death associated with variants in MYL3. Genet. Med. 2021, 23, 787–792. [Google Scholar] [CrossRef]
  91. Greenway, S.C.; Wilson, G.J.; Wilson, J.; George, K.; Kantor, P.F. Sudden death in an infant with angina, restrictive cardiomyopathy, and coronary artery bridging: An unusual phenotype for a beta-myosin heavy chain (MYH7) sarcomeric protein mutation. Circ. Heart Fail. 2012, 5, e92–e93. [Google Scholar] [CrossRef] [Green Version]
  92. Neagoe, O.; Ciobanu, A.; Diaconu, R.; Mirea, O.; Donoiu, I.; Militaru, C. A rare case of familial restrictive cardiomyopathy, with mutations in MYH7 and ABCC9 genes. Discoveries 2019, 7, e99. [Google Scholar] [CrossRef]
  93. Ware, S.M.; Quinn, M.E.; Ballard, E.T.; Miller, E.; Uzark, K.; Spicer, R.L. Pediatric restrictive cardiomyopathy associated with a mutation in beta-myosin heavy chain. Clin. Genet. 2008, 73, 165–170. [Google Scholar] [CrossRef]
  94. Kawano, H.; Kawamura, K.; Kanda, M.; Ishijima, M.; Abe, K.; Hayashi, T.; Matsumoto, Y.; Kimura, A.; Maemura, K. Histopathological changes of myocytes in restrictive cardiomyopathy. Med. Mol. Morphol. 2021, 54, 289–295. [Google Scholar] [CrossRef] [PubMed]
  95. Fan, L.L.; Guo, S.; Jin, J.Y.; He, Z.J.; Zhao, S.P.; Xiang, R.; Zhao, W. Whole exome sequencing identified a 13 base pair MYH7 deletion-mutation in a patient with restrictive cardiomyopathy and left ventricle hypertrophy. Ann. Clin. Lab. Sci. 2019, 49, 838–840. [Google Scholar] [PubMed]
  96. Gerull, B. The rapidly evolving role of titin in cardiac physiology and cardiomyopathy. Can. J. Cardiol. 2015, 31, 1351–1359. [Google Scholar] [CrossRef] [PubMed]
  97. Gerull, B.; Gramlich, M.; Atherton, J.; McNabb, M.; Trombitas, K.; Sasse-Klaassen, S.; Seidman, J.G.; Seidman, C.; Granzier, H.; Labeit, S.; et al. Mutations of TTN, encoding the giant muscle filament titin, cause familial dilated cardiomyopathy. Nat. Genet. 2002, 30, 201–204. [Google Scholar] [CrossRef]
  98. Herman, D.S.; Lam, L.; Taylor, M.R.; Wang, L.; Teekakirikul, P.; Christodoulou, D.; Conner, L.; DePalma, S.R.; McDonough, B.; Sparks, E.; et al. Truncations of titin causing dilated cardiomyopathy. N. Engl. J. Med. 2012, 366, 619–628. [Google Scholar] [CrossRef] [Green Version]
  99. Ware, J.S.; Cook, S.A. Role of titin in cardiomyopathy: From DNA variants to patient stratification. Nat. Rev. Cardiol. 2018, 15, 241–252. [Google Scholar] [CrossRef]
  100. McAfee, Q.; Chen, C.Y.; Yang, Y.; Caporizzo, M.A.; Morley, M.; Babu, A.; Jeong, S.; Brandimarto, J.; Bedi, K.C., Jr.; Flam, E.; et al. Truncated titin proteins in dilated cardiomyopathy. Sci. Transl. Med. 2021, 13, eabd7287. [Google Scholar] [CrossRef]
  101. Akinrinade, O.; Helio, T.; Lekanne Deprez, R.H.; Jongbloed, J.D.H.; Boven, L.G.; van den Berg, M.P.; Pinto, Y.M.; Alastalo, T.P.; Myllykangas, S.; Spaendonck-Zwarts, K.V.; et al. Relevance of titin missense and non-frameshifting insertions/deletions variants in dilated cardiomyopathy. Sci. Rep. 2019, 9, 4093. [Google Scholar] [CrossRef]
  102. Brodehl, A.; Gaertner-Rommel, A.; Milting, H. Molecular insights into cardiomyopathies associated with desmin (DES) mutations. Biophys. Rev. 2018, 10, 983–1006. [Google Scholar] [CrossRef]
  103. Dayal, A.A.; Medvedeva, N.V.; Nekrasova, T.M.; Duhalin, S.D.; Surin, A.K.; Minin, A.A. Desmin interacts directly with mitochondria. Int. J. Mol. Sci. 2020, 21, 8122. [Google Scholar] [CrossRef]
  104. Patel, D.M.; Green, K.J. Desmosomes in the heart: A review of clinical and mechanistic analyses. Cell Commun. Adhes. 2014, 21, 109–128. [Google Scholar] [CrossRef] [PubMed]
  105. Hatsell, S.; Cowin, P. Deconstructing desmoplakin. Nat. Cell Biol. 2001, 3, E270–E272. [Google Scholar] [CrossRef] [PubMed]
  106. Gorza, L.; Sorge, M.; Secli, L.; Brancaccio, M. Master Regulators of muscle atrophy: Role of costamere components. Cells 2021, 10, 61. [Google Scholar] [CrossRef] [PubMed]
  107. Wiche, G. Plectin-mediated intermediate filament functions: Why isoforms matter. Cells 2021, 10, 2154. [Google Scholar] [CrossRef] [PubMed]
  108. Li, Z.; Colucci-Guyon, E.; Pincon-Raymond, M.; Mericskay, M.; Pournin, S.; Paulin, D.; Babinet, C. Cardiovascular lesions and skeletal myopathy in mice lacking desmin. Dev. Biol. 1996, 175, 362–366. [Google Scholar] [CrossRef] [Green Version]
  109. Capetanaki, Y.; Milner, D.J.; Weitzer, G. Desmin in muscle formation and maintenance: Knockouts and consequences. Cell Struct. Funct. 1997, 22, 103–116. [Google Scholar] [CrossRef] [Green Version]
  110. Schirmer, I.; Dieding, M.; Klauke, B.; Brodehl, A.; Gaertner-Rommel, A.; Walhorn, V.; Gummert, J.; Schulz, U.; Paluszkiewicz, L.; Anselmetti, D.; et al. A novel desmin (DES) indel mutation causes severe atypical cardiomyopathy in combination with atrioventricular block and skeletal myopathy. Mol. Genet. Genom. Med. 2018, 6, 288–293. [Google Scholar] [CrossRef] [Green Version]
  111. Marakhonov, A.V.; Brodehl, A.; Myasnikov, R.P.; Sparber, P.A.; Kiseleva, A.V.; Kulikova, O.V.; Meshkov, A.N.; Zharikova, A.A.; Koretsky, S.N.; Kharlap, M.S.; et al. Noncompaction cardiomyopathy is caused by a novel in-frame desmin (DES) deletion mutation within the 1A coiled-coil rod segment leading to a severe filament assembly defect. Hum. Mutat. 2019, 40, 734–741. [Google Scholar] [CrossRef]
  112. Kubanek, M.; Schimerova, T.; Piherova, L.; Brodehl, A.; Krebsova, A.; Ratnavadivel, S.; Stanasiuk, C.; Hansikova, H.; Zeman, J.; Palecek, T.; et al. Desminopathy: Novel desmin variants, a new cardiac phenotype, and further evidence for secondary mitochondrial dysfunction. J. Clin. Med. 2020, 9, 937. [Google Scholar] [CrossRef] [Green Version]
  113. Protonotarios, A.; Brodehl, A.; Asimaki, A.; Jager, J.; Quinn, E.; Stanasiuk, C.; Ratnavadivel, S.; Futema, M.; Akhtar, M.M.; Gossios, T.D.; et al. The novel desmin variant p.Leu115Ile is associated with a unique form of biventricular arrhythmogenic cardiomyopathy. Can. J. Cardiol. 2021, 37, 857–866. [Google Scholar] [CrossRef]
  114. Fischer, B.; Dittmann, S.; Brodehl, A.; Unger, A.; Stallmeyer, B.; Paul, M.; Seebohm, G.; Kayser, A.; Peischard, S.; Linke, W.A.; et al. Functional characterization of novel alpha-helical rod domain desmin (DES) pathogenic variants associated with dilated cardiomyopathy, atrioventricular block and a risk for sudden cardiac death. Int. J. Cardiol. 2021, 329, 167–174. [Google Scholar] [CrossRef] [PubMed]
  115. Clemen, C.S.; Fischer, D.; Reimann, J.; Eichinger, L.; Muller, C.R.; Muller, H.D.; Goebel, H.H.; Schroder, R. How much mutant protein is needed to cause a protein aggregate myopathy in vivo? Lessons from an exceptional desminopathy. Hum. Mutat. 2009, 30, E490–E499. [Google Scholar] [CrossRef] [PubMed]
  116. Brodehl, A.; Hain, C.; Flottmann, F.; Ratnavadivel, S.; Gaertner, A.; Klauke, B.; Kalinowski, J.; Körperich, H.; Gummert, J.; Paluszkiewicz, L.; et al. The desmin mutation DES-c.735G>C causes severe restrictive cardiomyopathy by inducing in-frame skipping of exon-3. Biomedicines 2021, 9, 1400. [Google Scholar] [CrossRef] [PubMed]
  117. Chen, Z.; Li, R.; Wang, Y.; Cao, L.; Lin, C.; Liu, F.; Hu, R.; Nan, J.; Zhuang, X.; Lu, X.; et al. Features of myocardial injury detected by cardiac magnetic resonance in a patient with desmin-related restrictive cardiomyopathy. ESC Heart Fail. 2021, 8, 5560–5564. [Google Scholar] [CrossRef]
  118. Herrmann, H.; Cabet, E.; Chevalier, N.R.; Moosmann, J.; Schultheis, D.; Haas, J.; Schowalter, M.; Berwanger, C.; Weyerer, V.; Agaimy, A.; et al. Dual Functional states of R406W-desmin assembly complexes cause cardiomyopathy with severe intercalated disc derangement in humans and in knock-in mice. Circulation 2020, 142, 2155–2171. [Google Scholar] [CrossRef]
  119. Ojrzynska, N.; Bilinska, Z.T.; Franaszczyk, M.; Ploski, R.; Grzybowski, J. Restrictive cardiomyopathy due to novel desmin gene mutation. Kardiol. Pol. 2017, 75, 723. [Google Scholar] [CrossRef] [Green Version]
  120. Jurcu, T.R.; Bastian, A.E.; Militaru, S.; Popa, A.; Manole, E.; Popescu, B.A.; Tallila, J.; Popescu, B.O.; Ginghina, C.D. Discovery of a new mutation in the desmin gene in a young patient with cardiomyopathy and muscular weakness. Rom. J. Morphol. Embryol. 2017, 58, 225–230. [Google Scholar]
  121. Sharma, S.; Juneja, R.; Sharma, G.; Arava, S.; Ray, R. Desmin-related restrictive cardiomyopathy in a pediatric patient: A case report. Indian J. Pathol. Microbiol. 2013, 56, 402–404. [Google Scholar]
  122. Pinol-Ripoll, G.; Shatunov, A.; Cabello, A.; Larrode, P.; de la Puerta, I.; Pelegrin, J.; Ramos, F.J.; Olive, M.; Goldfarb, L.G. Severe infantile-onset cardiomyopathy associated with a homozygous deletion in desmin. Neuromuscul. Disord. 2009, 19, 418–422. [Google Scholar] [CrossRef] [Green Version]
  123. Bar, H.; Mucke, N.; Kostareva, A.; Sjoberg, G.; Aebi, U.; Herrmann, H. Severe muscle disease-causing desmin mutations interfere with in vitro filament assembly at distinct stages. Proc. Natl. Acad. Sci. USA 2005, 102, 15099–15104. [Google Scholar] [CrossRef] [Green Version]
  124. Brodehl, A.; Hedde, P.N.; Dieding, M.; Fatima, A.; Walhorn, V.; Gayda, S.; Saric, T.; Klauke, B.; Gummert, J.; Anselmetti, D.; et al. Dual color photoactivation localization microscopy of cardiomyopathy-associated desmin mutants. J. Biol. Chem. 2012, 287, 16047–16057. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Herrmann, H.; Aebi, U. Intermediate filaments: Structure and assembly. Cold Spring Harb. Perspect. Biol. 2016, 8, a018242. [Google Scholar] [CrossRef] [PubMed]
  126. Quinlan, R.A.; Hatzfeld, M.; Franke, W.W.; Lustig, A.; Schulthess, T.; Engel, J. Characterization of dimer subunits of intermediate filament proteins. J. Mol. Biol. 1986, 192, 337–349. [Google Scholar] [CrossRef]
  127. Parry, D.A.; Strelkov, S.V.; Burkhard, P.; Aebi, U.; Herrmann, H. Towards a molecular description of intermediate filament structure and assembly. Exp. Cell Res. 2007, 313, 2204–2216. [Google Scholar] [CrossRef]
  128. Herrmann, H.; Haner, M.; Brettel, M.; Ku, N.O.; Aebi, U. Characterization of distinct early assembly units of different intermediate filament proteins. J. Mol. Biol. 1999, 286, 1403–1420. [Google Scholar] [CrossRef]
  129. Herrmann, H.; Aebi, U. Intermediate filaments: Molecular structure, assembly mechanism, and integration into functionally distinct intracellular Scaffolds. Annu. Rev. Biochem. 2004, 73, 749–789. [Google Scholar] [CrossRef]
  130. Colakoglu, G.; Brown, A. Intermediate filaments exchange subunits along their length and elongate by end-to-end annealing. J. Cell Biol. 2009, 185, 769–777. [Google Scholar] [CrossRef] [Green Version]
  131. Winheim, S.; Hieb, A.R.; Silbermann, M.; Surmann, E.M.; Wedig, T.; Herrmann, H.; Langowski, J.; Mucke, N. Deconstructing the late phase of vimentin assembly by total internal reflection fluorescence microscopy (TIRFM). PLoS ONE 2011, 6, e19202. [Google Scholar] [CrossRef] [Green Version]
  132. Noding, B.; Herrmann, H.; Koster, S. Direct observation of subunit exchange along mature vimentin intermediate filaments. Biophys. J. 2014, 107, 2923–2931. [Google Scholar] [CrossRef] [Green Version]
  133. Park, K.Y.; Dalakas, M.C.; Goebel, H.H.; Ferrans, V.J.; Semino-Mora, C.; Litvak, S.; Takeda, K.; Goldfarb, L.G. Desmin splice variants causing cardiac and skeletal myopathy. J. Med. Genet. 2000, 37, 851–857. [Google Scholar] [CrossRef]
  134. Arbustini, E.; Pasotti, M.; Pilotto, A.; Pellegrini, C.; Grasso, M.; Previtali, S.; Repetto, A.; Bellini, O.; Azan, G.; Scaffino, M.; et al. Desmin accumulation restrictive cardiomyopathy and atrioventricular block associated with desmin gene defects. Eur. J. Heart Fail. 2006, 8, 477–483. [Google Scholar] [CrossRef] [PubMed]
  135. Olive, M.; Armstrong, J.; Miralles, F.; Pou, A.; Fardeau, M.; Gonzalez, L.; Martinez, F.; Fischer, D.; Martinez Matos, J.A.; Shatunov, A.; et al. Phenotypic patterns of desminopathy associated with three novel mutations in the desmin gene. Neuromuscul. Disord. 2007, 17, 443–450. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Pruszczyk, P.; Kostera-Pruszczyk, A.; Shatunov, A.; Goudeau, B.; Draminska, A.; Takeda, K.; Sambuughin, N.; Vicart, P.; Strelkov, S.V.; Goldfarb, L.G.; et al. Restrictive cardiomyopathy with atrioventricular conduction block resulting from a desmin mutation. Int. J. Cardiol. 2007, 117, 244–253. [Google Scholar] [CrossRef] [PubMed]
  137. Bar, H.; Goudeau, B.; Walde, S.; Casteras-Simon, M.; Mucke, N.; Shatunov, A.; Goldberg, Y.P.; Clarke, C.; Holton, J.L.; Eymard, B.; et al. Conspicuous involvement of desmin tail mutations in diverse cardiac and skeletal myopathies. Hum. Mutat. 2007, 28, 374–386. [Google Scholar] [CrossRef] [PubMed]
  138. Bang, M.L.; Mudry, R.E.; McElhinny, A.S.; Trombitas, K.; Geach, A.J.; Yamasaki, R.; Sorimachi, H.; Granzier, H.; Gregorio, C.C.; Labeit, S. Myopalladin, a novel 145-kilodalton sarcomeric protein with multiple roles in Z-disc and I-band protein assemblies. J. Cell Biol. 2001, 153, 413–427. [Google Scholar] [CrossRef] [Green Version]
  139. Filomena, M.C.; Yamamoto, D.L.; Carullo, P.; Medvedev, R.; Ghisleni, A.; Piroddi, N.; Scellini, B.; Crispino, R.; D’Autilia, F.; Zhang, J.; et al. Myopalladin knockout mice develop cardiac dilation and show a maladaptive response to mechanical pressure overload. eLife 2021, 10, e58313. [Google Scholar] [CrossRef]
  140. Duboscq-Bidot, L.; Xu, P.; Charron, P.; Neyroud, N.; Dilanian, G.; Millaire, A.; Bors, V.; Komajda, M.; Villard, E. Mutations in the Z-band protein myopalladin gene and idiopathic dilated cardiomyopathy. Cardiovasc. Res. 2008, 77, 118–125. [Google Scholar] [CrossRef]
  141. Bagnall, R.D.; Yeates, L.; Semsarian, C. Analysis of the Z-disc genes PDLIM3 and MYPN in patients with hypertrophic cardiomyopathy. Int. J. Cardiol. 2010, 145, 601–602. [Google Scholar] [CrossRef]
  142. Miyatake, S.; Mitsuhashi, S.; Hayashi, Y.K.; Purevjav, E.; Nishikawa, A.; Koshimizu, E.; Suzuki, M.; Yatabe, K.; Tanaka, Y.; Ogata, K.; et al. Biallelic mutations in MYPN, encoding myopalladin, are associated with childhood-onset, slowly progressive nemaline myopathy. Am. J. Hum. Genet. 2017, 100, 169–178. [Google Scholar] [CrossRef] [Green Version]
  143. Beggs, A.H.; Byers, T.J.; Knoll, J.H.; Boyce, F.M.; Bruns, G.A.; Kunkel, L.M. Cloning and characterization of two human skeletal muscle alpha-actinin genes located on chromosomes 1 and 11. J. Biol. Chem. 1992, 267, 9281–9288. [Google Scholar] [CrossRef]
  144. Tiso, N.; Majetti, M.; Stanchi, F.; Rampazzo, A.; Zimbello, R.; Nava, A.; Danieli, G.A. Fine mapping and genomic structure of ACTN2, the human gene coding for the sarcomeric isoform of α-actinin-2, expressed in skeletal and cardiac muscle. Biochem. Biophys. Res. Commun. 1999, 265, 256–259. [Google Scholar] [CrossRef] [PubMed]
  145. Wadmore, K.; Azad, A.J.; Gehmlich, K. The role of Z-disc proteins in myopathy and cardiomyopathy. Int. J. Mol. Sci. 2021, 22, 3058. [Google Scholar] [CrossRef] [PubMed]
  146. Nicolas, A.; Delalande, O.; Hubert, J.F.; Le Rumeur, E. The spectrin family of proteins: A unique coiled-coil fold for various molecular surface properties. J. Struct. Biol. 2014, 186, 392–401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Davison, M.D.; Critchley, D.R. alpha-Actinins and the DMD protein contain spectrin-like repeats. Cell 1988, 52, 159–160. [Google Scholar] [CrossRef]
  148. Djinovic-Carugo, K.; Gautel, M.; Ylanne, J.; Young, P. The spectrin repeat: A structural platform for cytoskeletal protein assemblies. FEBS Lett. 2002, 513, 119–123. [Google Scholar] [CrossRef] [Green Version]
  149. Ribeiro Ede, A., Jr.; Pinotsis, N.; Ghisleni, A.; Salmazo, A.; Konarev, P.V.; Kostan, J.; Sjoblom, B.; Schreiner, C.; Polyansky, A.A.; Gkougkoulia, E.A.; et al. The structure and regulation of human muscle α-actinin. Cell 2014, 159, 1447–1460. [Google Scholar] [CrossRef] [Green Version]
  150. Mohapatra, B.; Jimenez, S.; Lin, J.H.; Bowles, K.R.; Coveler, K.J.; Marx, J.G.; Chrisco, M.A.; Murphy, R.T.; Lurie, P.R.; Schwartz, R.J.; et al. Mutations in the muscle LIM protein and α-actinin-2 genes in dilated cardiomyopathy and endocardial fibroelastosis. Mol. Genet. Metab. 2003, 80, 207–215. [Google Scholar] [CrossRef]
  151. Prondzynski, M.; Lemoine, M.D.; Zech, A.T.; Horvath, A.; Di Mauro, V.; Koivumaki, J.T.; Kresin, N.; Busch, J.; Krause, T.; Kramer, E.; et al. Disease modeling of a mutation in α-actinin 2 guides clinical therapy in hypertrophic cardiomyopathy. EMBO Mol. Med. 2019, 11, e11115. [Google Scholar] [CrossRef]
  152. Park, J.; Cho, Y.G.; Park, H.W.; Cho, J.S. Case report: Novel likely pathogenic ACTN2 variant causing heterogeneous phenotype in a korean family with left ventricular non-compaction. Front. Pediatr. 2021, 9, 609389. [Google Scholar] [CrossRef]
  153. Good, J.M.; Fellmann, F.; Bhuiyan, Z.A.; Rotman, S.; Pruvot, E.; Schlapfer, J. ACTN2 variant associated with a cardiac phenotype suggestive of left-dominant arrhythmogenic cardiomyopathy. HeartRhythm Case Rep. 2020, 6, 15–19. [Google Scholar] [CrossRef] [Green Version]
  154. Inoue, M.; Noguchi, S.; Sonehara, K.; Nakamura-Shindo, K.; Taniguchi, A.; Kajikawa, H.; Nakamura, H.; Ishikawa, K.; Ogawa, M.; Hayashi, S.; et al. A recurrent homozygous ACTN2 variant associated with core myopathy. Acta Neuropathol. 2021, 142, 785–788. [Google Scholar] [CrossRef] [PubMed]
  155. Vorgerd, M.; van der Ven, P.F.; Bruchertseifer, V.; Lowe, T.; Kley, R.A.; Schroder, R.; Lochmuller, H.; Himmel, M.; Koehler, K.; Furst, D.O.; et al. A mutation in the dimerization domain of filamin c causes a novel type of autosomal dominant myofibrillar myopathy. Am. J. Hum. Genet. 2005, 77, 297–304. [Google Scholar] [CrossRef] [Green Version]
  156. Duff, R.M.; Tay, V.; Hackman, P.; Ravenscroft, G.; McLean, C.; Kennedy, P.; Steinbach, A.; Schoffler, W.; van der Ven, P.F.M.; Furst, D.O.; et al. Mutations in the N-terminal actin-binding domain of filamin C cause a distal myopathy. Am. J. Hum. Genet. 2011, 88, 729–740. [Google Scholar] [CrossRef] [Green Version]
  157. Chakarova, C.; Wehnert, M.S.; Uhl, K.; Sakthivel, S.; Vosberg, H.P.; van der Ven, P.F.; Furst, D.O. Genomic structure and fine mapping of the two human filamin gene paralogues FLNB and FLNC and comparative analysis of the filamin gene family. Hum. Genet. 2000, 107, 597–611. [Google Scholar] [CrossRef] [PubMed]
  158. Mao, Z.; Nakamura, F. Structure and function of filamin c in the muscle Z-disc. Int. J. Mol. Sci. 2020, 21, 2696. [Google Scholar] [CrossRef] [Green Version]
  159. Pudas, R.; Kiema, T.R.; Butler, P.J.; Stewart, M.; Ylanne, J. Structural basis for vertebrate filamin dimerization. Structure 2005, 13, 111–119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. van der Ven, P.F.; Obermann, W.M.; Lemke, B.; Gautel, M.; Weber, K.; Furst, D.O. Characterization of muscle filamin isoforms suggests a possible role of gamma-filamin/ABP-L in sarcomeric Z-disc formation. Cell Motil. Cytoskelet. 2000, 45, 149–162. [Google Scholar] [CrossRef]
  161. Labeit, S.; Lahmers, S.; Burkart, C.; Fong, C.; McNabb, M.; Witt, S.; Witt, C.; Labeit, D.; Granzier, H. Expression of distinct classes of titin isoforms in striated and smooth muscles by alternative splicing, and their conserved interaction with filamins. J. Mol. Biol. 2006, 362, 664–681. [Google Scholar] [CrossRef]
  162. Gonzalez-Morales, N.; Holenka, T.K.; Schock, F. Filamin actin-binding and titin-binding fulfill distinct functions in Z-disc cohesion. PLoS Genet. 2017, 13, e1006880. [Google Scholar] [CrossRef]
  163. Gontier, Y.; Taivainen, A.; Fontao, L.; Sonnenberg, A.; van der Flier, A.; Carpen, O.; Faulkner, G.; Borradori, L. The Z-disc proteins myotilin and FATZ-1 interact with each other and are connected to the sarcolemma via muscle-specific filamins. J. Cell Sci. 2005, 118 Pt 16, 3739–3749. [Google Scholar] [CrossRef] [Green Version]
  164. Thompson, T.G.; Chan, Y.M.; Hack, A.A.; Brosius, M.; Rajala, M.; Lidov, H.G.; McNally, E.M.; Watkins, S.; Kunkel, L.M. Filamin 2 (FLN2): A muscle-specific sarcoglycan interacting protein. J. Cell Biol. 2000, 148, 115–126. [Google Scholar] [CrossRef] [PubMed]
  165. Valdes-Mas, R.; Gutierrez-Fernandez, A.; Gomez, J.; Coto, E.; Astudillo, A.; Puente, D.A.; Reguero, J.R.; Alvarez, V.; Moris, C.; Leon, D.; et al. Mutations in filamin C cause a new form of familial hypertrophic cardiomyopathy. Nat. Commun. 2014, 5, 5326. [Google Scholar] [CrossRef] [PubMed]
  166. Begay, R.L.; Tharp, C.A.; Martin, A.; Graw, S.L.; Sinagra, G.; Miani, D.; Sweet, M.E.; Slavov, D.B.; Stafford, N.; Zeller, M.J.; et al. FLNC gene splice mutations cause dilated cardiomyopathy. JACC Basic Transl. Sci. 2016, 1, 344–359. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Augusto, J.B.; Eiros, R.; Nakou, E.; Moura-Ferreira, S.; Treibel, T.A.; Captur, G.; Akhtar, M.M.; Protonotarios, A.; Gossios, T.D.; Savvatis, K.; et al. Dilated cardiomyopathy and arrhythmogenic left ventricular cardiomyopathy: A comprehensive genotype-imaging phenotype study. Eur. Heart J. Cardiovasc. Imaging 2020, 21, 326–336. [Google Scholar] [CrossRef]
  168. Van Waning, J.I.; Hoedemaekers, Y.M.; te Rijdt, W.P.; Jpma, A.I.; Heijsman, D.; Caliskan, K.; Hoendermis, E.S.; Willems, T.P.; van den Wijngaard, A.; Suurmeijer, A. FLNC missense variants in familial noncompaction cardiomyopathy. Cardiogenetics 2019, 9, 9–13. [Google Scholar] [CrossRef] [Green Version]
  169. Reinstein, E.; Gutierrez-Fernandez, A.; Tzur, S.; Bormans, C.; Marcu, S.; Tayeb-Fligelman, E.; Vinkler, C.; Raas-Rothschild, A.; Irge, D.; Landau, M.; et al. Congenital dilated cardiomyopathy caused by biallelic mutations in Filamin C. Eur. J. Hum. Genet. 2016, 24, 1792–1796. [Google Scholar] [CrossRef] [Green Version]
  170. Chen, J.; Wu, J.; Han, C.; Li, Y.; Guo, Y.; Tong, X. A mutation in the filamin c gene causes myofibrillar myopathy with lower motor neuron syndrome: A case report. BMC Neurol. 2019, 19, 198. [Google Scholar] [CrossRef] [Green Version]
  171. Dalkilic, I.; Schienda, J.; Thompson, T.G.; Kunkel, L.M. Loss of FilaminC (FLNc) results in severe defects in myogenesis and myotube structure. Mol. Cell. Biol. 2006, 26, 6522–6534. [Google Scholar] [CrossRef] [Green Version]
  172. Ruparelia, A.A.; Zhao, M.; Currie, P.D.; Bryson-Richardson, R.J. Characterization and investigation of zebrafish models of filamin-related myofibrillar myopathy. Hum. Mol. Genet. 2012, 21, 4073–4083. [Google Scholar] [CrossRef] [Green Version]
  173. Fujita, M.; Mitsuhashi, H.; Isogai, S.; Nakata, T.; Kawakami, A.; Nonaka, I.; Noguchi, S.; Hayashi, Y.K.; Nishino, I.; Kudo, A. Filamin C plays an essential role in the maintenance of the structural integrity of cardiac and skeletal muscles, revealed by the medaka mutant zacro. Dev. Biol. 2012, 361, 79–89. [Google Scholar] [CrossRef] [Green Version]
  174. Deo, R.C.; Musso, G.; Tasan, M.; Tang, P.; Poon, A.; Yuan, C.; Felix, J.F.; Vasan, R.S.; Beroukhim, R.; De Marco, T.; et al. Prioritizing causal disease genes using unbiased genomic features. Genome Biol. 2014, 15, 534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Chevessier, F.; Schuld, J.; Orfanos, Z.; Plank, A.C.; Wolf, L.; Maerkens, A.; Unger, A.; Schlotzer-Schrehardt, U.; Kley, R.A.; Von Horsten, S.; et al. Myofibrillar instability exacerbated by acute exercise in filaminopathy. Hum. Mol. Genet. 2015, 24, 7207–7220. [Google Scholar] [CrossRef] [PubMed]
  176. Kiselev, A.; Vaz, R.; Knyazeva, A.; Khudiakov, A.; Tarnovskaya, S.; Liu, J.; Sergushichev, A.; Kazakov, S.; Frishman, D.; Smolina, N.; et al. De novo mutations in FLNC leading to early-onset restrictive cardiomyopathy and congenital myopathy. Hum. Mutat. 2018, 39, 1161–1172. [Google Scholar] [CrossRef] [PubMed]
  177. Ruparelia, A.A.; Oorschot, V.; Ramm, G.; Bryson-Richardson, R.J. FLNC myofibrillar myopathy results from impaired autophagy and protein insufficiency. Hum. Mol. Genet. 2016, 25, 2131–2142. [Google Scholar] [CrossRef]
  178. Zhou, Y.; Chen, Z.; Zhang, L.; Zhu, M.; Tan, C.; Zhou, X.; Evans, S.M.; Fang, X.; Feng, W.; Chen, J. Loss of Filamin C is catastrophic for heart function. Circulation 2020, 141, 869–871. [Google Scholar] [CrossRef]
  179. Rodina, N.; Khudiakov, A.; Perepelina, K.; Muravyev, A.; Boytsov, A.; Zlotina, A.; Sokolnikova, P.; Kostareva, A. Generation of iPSC line (FAMRCi009-A) from patient with familial progressive cardiac conduction disorder carrying genetic variant FLNC p.Val2264Met. Stem Cell Res. 2021, 59, 102640. [Google Scholar] [CrossRef]
  180. Perepelina, K.; Khudiakov, A.; Rodina, N.; Boytsov, A.; Vavilova, T.; Zlotina, A.; Sokolnikova, P.; Kostareva, A. Generation of iPSC line FAMRCi010-A from patient with restrictive cardiomyopathy carrying genetic variant FLNC p.Gly2011Arg. Stem Cell Res. 2021, 59, 102639. [Google Scholar] [CrossRef]
  181. Tucker, N.R.; McLellan, M.A.; Hu, D.; Ye, J.; Parsons, V.A.; Mills, R.W.; Clauss, S.; Dolmatova, E.; Shea, M.A.; Milan, D.J.; et al. Novel mutation in FLNC (Filamin C) causes familial restrictive cardiomyopathy. Circ. Cardiovasc. Genet. 2017, 10, e001780. [Google Scholar] [CrossRef] [Green Version]
  182. Xiao, F.; Wei, Q.; Wu, B.; Liu, X.; Mading, A.; Yang, L.; Li, Y.; Liu, F.; Pan, X.; Wang, H. Clinical exome sequencing revealed that FLNC variants contribute to the early diagnosis of cardiomyopathies in infant patients. Transl. Pediatr. 2020, 9, 21–33. [Google Scholar] [CrossRef]
  183. Roldan-Sevilla, A.; Palomino-Doza, J.; de Juan, J.; Sanchez, V.; Dominguez-Gonzalez, C.; Salguero-Bodes, R.; Arribas-Ynsaurriaga, F. Missense mutations in the FLNC gene causing familial restrictive cardiomyopathy. Circ. Genom. Precis. Med. 2019, 12, e002388. [Google Scholar] [CrossRef]
  184. Schubert, J.; Tariq, M.; Geddes, G.; Kindel, S.; Miller, E.M.; Ware, S.M. Novel pathogenic variants in Filamin C identified in pediatric restrictive cardiomyopathy. Hum. Mutat. 2018, 39, 2083–2096. [Google Scholar] [CrossRef] [PubMed]
  185. Aebi, U.; Cohn, J.; Buhle, L.; Gerace, L. The nuclear lamina is a meshwork of intermediate-type filaments. Nature 1986, 323, 560–564. [Google Scholar] [CrossRef] [PubMed]
  186. Dobrzynska, A.; Gonzalo, S.; Shanahan, C.; Askjaer, P. The nuclear lamina in health and disease. Nucleus 2016, 7, 233–248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Fatkin, D.; MacRae, C.; Sasaki, T.; Wolff, M.R.; Porcu, M.; Frenneaux, M.; Atherton, J.; Vidaillet, H.J., Jr.; Spudich, S.; De Girolami, U.; et al. Missense mutations in the rod domain of the lamin A/C gene as causes of dilated cardiomyopathy and conduction-system disease. N. Engl. J. Med. 1999, 341, 1715–1724. [Google Scholar] [CrossRef] [Green Version]
  188. Quarta, G.; Syrris, P.; Ashworth, M.; Jenkins, S.; Zuborne Alapi, K.; Morgan, J.; Muir, A.; Pantazis, A.; McKenna, W.J.; Elliott, P.M. Mutations in the Lamin A/C gene mimic arrhythmogenic right ventricular cardiomyopathy. Eur. Heart J. 2012, 33, 1128–1136. [Google Scholar] [CrossRef] [Green Version]
  189. Liu, Z.; Shan, H.; Huang, J.; Li, N.; Hou, C.; Pu, J. A novel lamin A/C gene missense mutation (445 V > E) in immunoglobulin-like fold associated with left ventricular non-compaction. Europace 2016, 18, 617–622. [Google Scholar] [CrossRef]
  190. Raffaele Di Barletta, M.; Ricci, E.; Galluzzi, G.; Tonali, P.; Mora, M.; Morandi, L.; Romorini, A.; Voit, T.; Orstavik, K.H.; Merlini, L.; et al. Different mutations in the LMNA gene cause autosomal dominant and autosomal recessive Emery-Dreifuss muscular dystrophy. Am. J. Hum. Genet. 2000, 66, 1407–1412. [Google Scholar] [CrossRef] [Green Version]
  191. Shackleton, S.; Lloyd, D.J.; Jackson, S.N.; Evans, R.; Niermeijer, M.F.; Singh, B.M.; Schmidt, H.; Brabant, G.; Kumar, S.; Durrington, P.N.; et al. LMNA, encoding lamin A/C, is mutated in partial lipodystrophy. Nat. Genet. 2000, 24, 153–156. [Google Scholar] [CrossRef]
  192. Eriksson, M.; Brown, W.T.; Gordon, L.B.; Glynn, M.W.; Singer, J.; Scott, L.; Erdos, M.R.; Robbins, C.M.; Moses, T.Y.; Berglund, P.; et al. Recurrent de novo point mutations in lamin A cause Hutchinson-Gilford progeria syndrome. Nature 2003, 423, 293–298. [Google Scholar] [CrossRef] [Green Version]
  193. Merner, N.D.; Hodgkinson, K.A.; Haywood, A.F.; Connors, S.; French, V.M.; Drenckhahn, J.D.; Kupprion, C.; Ramadanova, K.; Thierfelder, L.; McKenna, W.; et al. Arrhythmogenic right ventricular cardiomyopathy type 5 is a fully penetrant, lethal arrhythmic disorder caused by a missense mutation in the TMEM43 gene. Am. J. Hum. Genet. 2008, 82, 809–821. [Google Scholar] [CrossRef] [Green Version]
  194. Hirata, T.; Fujita, M.; Nakamura, S.; Gotoh, K.; Motooka, D.; Murakami, Y.; Maeda, Y.; Kinoshita, T. Post-Golgi anterograde transport requires GARP-dependent endosome-to-TGN retrograde transport. Mol. Biol. Cell 2015, 26, 3071–3084. [Google Scholar] [CrossRef] [PubMed]
  195. Russell, M.W.; Raeker, M.O.; Geisler, S.B.; Thomas, P.E.; Simmons, T.A.; Bernat, J.A.; Thorsson, T.; Innis, J.W. Functional analysis of candidate genes in 2q13 deletion syndrome implicates FBLN7 and TMEM87B deficiency in congenital heart defects and FBLN7 in craniofacial malformations. Hum. Mol. Genet. 2014, 23, 4272–4284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Garrido, C.; Paul, C.; Seigneuric, R.; Kampinga, H.H. The small heat shock proteins family: The long forgotten chaperones. Int. J. Biochem. Cell Biol. 2012, 44, 1588–1592. [Google Scholar] [CrossRef] [PubMed]
  197. Dubin, R.A.; Ally, A.H.; Chung, S.; Piatigorsky, J. Human αB-crystallin gene and preferential promoter function in lens. Genomics 1990, 7, 594–601. [Google Scholar] [CrossRef]
  198. Dimauro, I.; Antonioni, A.; Mercatelli, N.; Caporossi, D. The role of αB-crystallin in skeletal and cardiac muscle tissues. Cell Stress Chaperones 2018, 23, 491–505. [Google Scholar] [CrossRef]
  199. Chepelinsky, A.B.; Piatigorsky, J.; Pisano, M.M.; Dubin, R.A.; Wistow, G.; Limjoco, T.I.; Klement, J.F.; Jaworski, C.J. Lens protein gene expression: Alpha-crystallins and MIP. Lens Eye Toxic Res. 1991, 8, 319–344. [Google Scholar]
  200. Sacconi, S.; Feasson, L.; Antoine, J.C.; Pecheux, C.; Bernard, R.; Cobo, A.M.; Casarin, A.; Salviati, L.; Desnuelle, C.; Urtizberea, A. A novel CRYAB mutation resulting in multisystemic disease. Neuromuscul. Disord. 2012, 22, 66–72. [Google Scholar] [CrossRef]
  201. Safieh, L.A.; Khan, A.O.; Alkuraya, F.S. Identification of a novel CRYAB mutation associated with autosomal recessive juvenile cataract in a Saudi family. Mol. Vis. 2009, 15, 980–984. [Google Scholar]
  202. Inagaki, N.; Hayashi, T.; Arimura, T.; Koga, Y.; Takahashi, M.; Shibata, H.; Teraoka, K.; Chikamori, T.; Yamashina, A.; Kimura, A. αB-crystallin mutation in dilated cardiomyopathy. Biochem. Biophys. Res. Commun. 2006, 342, 379–386. [Google Scholar] [CrossRef]
  203. Peschek, J.; Braun, N.; Franzmann, T.M.; Georgalis, Y.; Haslbeck, M.; Weinkauf, S.; Buchner, J. The eye lens chaperone α-crystallin forms defined globular assemblies. Proc. Natl. Acad. Sci. USA 2009, 106, 13272–13277. [Google Scholar] [CrossRef] [Green Version]
  204. Ganea, E. Chaperone-like activity of alpha-crystallin and other small heat shock proteins. Curr. Protein Pept. Sci. 2001, 2, 205–225. [Google Scholar] [CrossRef] [PubMed]
  205. Bullard, B.; Ferguson, C.; Minajeva, A.; Leake, M.C.; Gautel, M.; Labeit, D.; Ding, L.; Labeit, S.; Horwitz, J.; Leonard, K.R.; et al. Association of the chaperone αB-crystallin with titin in heart muscle. J. Biol. Chem. 2004, 279, 7917–7924. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Jehle, S.; Rajagopal, P.; Bardiaux, B.; Markovic, S.; Kuhne, R.; Stout, J.R.; Higman, V.A.; Klevit, R.E.; van Rossum, B.J.; Oschkinat, H. Solid-state NMR and SAXS studies provide a structural basis for the activation of αB-crystallin oligomers. Nat. Struct. Mol. Biol. 2010, 17, 1037–1042. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Rajasekaran, N.S.; Connell, P.; Christians, E.S.; Yan, L.J.; Taylor, R.P.; Orosz, A.; Zhang, X.Q.; Stevenson, T.J.; Peshock, R.M.; Leopold, J.A.; et al. Human αB-crystallin mutation causes oxido-reductive stress and protein aggregation cardiomyopathy in mice. Cell 2007, 130, 427–439. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Sturner, E.; Behl, C. The role of the multifunctional BAG3 protein in cellular protein quality control and in disease. Front. Mol. Neurosci. 2017, 10, 177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Takayama, S.; Xie, Z.; Reed, J.C. An evolutionarily conserved family of Hsp70/Hsc70 molecular chaperone regulators. J. Biol. Chem. 1999, 274, 781–786. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Lin, H.; Koren, S.A.; Cvetojevic, G.; Girardi, P.; Johnson, G.V.W. The role of BAG3 in health and disease: A “Magic BAG of Tricks”. J. Cell. Biochem. 2022, 123, 4–21. [Google Scholar] [CrossRef]
  211. Kogel, D.; Linder, B.; Brunschweiger, A.; Chines, S.; Behl, C. At the crossroads of apoptosis and autophagy: Multiple roles of the Co-chaperone bag3 in stress and therapy resistance of cancer. Cells 2020, 9, 574. [Google Scholar] [CrossRef] [Green Version]
  212. Takayama, S.; Reed, J.C. Molecular chaperone targeting and regulation by BAG family proteins. Nat. Cell Biol. 2001, 3, E237–E241. [Google Scholar] [CrossRef]
  213. Sondermann, H.; Scheufler, C.; Schneider, C.; Hohfeld, J.; Hartl, F.U.; Moarefi, I. Structure of a Bag/Hsc70 complex: Convergent functional evolution of Hsp70 nucleotide exchange factors. Science 2001, 291, 1553–1557. [Google Scholar] [CrossRef]
  214. Shemetov, A.A.; Gusev, N.B. Biochemical characterization of small heat shock protein HspB8 (Hsp22)-Bag3 interaction. Arch. Biochem. Biophys. 2011, 513, 1–9. [Google Scholar] [CrossRef] [PubMed]
  215. Morelli, F.F.; Mediani, L.; Heldens, L.; Bertacchini, J.; Bigi, I.; Carra, A.D.; Vinet, J.; Carra, S. An interaction study in mammalian cells demonstrates weak binding of HSPB2 to BAG3, which is regulated by HSPB3 and abrogated by HSPB8. Cell Stress Chaperones 2017, 22, 531–540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Hishiya, A.; Salman, M.N.; Carra, S.; Kampinga, H.H.; Takayama, S. BAG3 directly interacts with mutated αB-crystallin to suppress its aggregation and toxicity. PLoS ONE 2011, 6, e16828. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Rauch, J.N.; Tse, E.; Freilich, R.; Mok, S.A.; Makley, L.N.; Southworth, D.R.; Gestwicki, J.E. BAG3 is a modular, scaffolding protein that physically links heat shock protein 70 (Hsp70) to the small heat shock proteins. J. Mol. Biol. 2017, 429, 128–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Fuchs, M.; Poirier, D.J.; Seguin, S.J.; Lambert, H.; Carra, S.; Charette, S.J.; Landry, J. Identification of the key structural motifs involved in HspB8/HspB6-Bag3 interaction. Biochem. J. 2009, 425, 245–255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Mizushima, W.; Sadoshima, J. BAG3 plays a central role in proteostasis in the heart. J. Clin. Investig. 2017, 127, 2900–2903. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. Norton, N.; Li, D.; Rieder, M.J.; Siegfried, J.D.; Rampersaud, E.; Zuchner, S.; Mangos, S.; Gonzalez-Quintana, J.; Wang, L.; McGee, S.; et al. Genome-wide studies of copy number variation and exome sequencing identify rare variants in BAG3 as a cause of dilated cardiomyopathy. Am. J. Hum. Genet. 2011, 88, 273–282. [Google Scholar] [CrossRef] [Green Version]
  221. Semmler, A.L.; Sacconi, S.; Bach, J.E.; Liebe, C.; Burmann, J.; Kley, R.A.; Ferbert, A.; Anderheiden, R.; Van den Bergh, P.; Martin, J.J.; et al. Unusual multisystemic involvement and a novel BAG3 mutation revealed by NGS screening in a large cohort of myofibrillar myopathies. Orphanet J. Rare Dis. 2014, 9, 121. [Google Scholar] [CrossRef] [Green Version]
  222. Kimura, K.; Ooms, A.; Graf-Riesen, K.; Kuppusamy, M.; Unger, A.; Schuld, J.; Daerr, J.; Lother, A.; Geisen, C.; Hein, L.; et al. Overexpression of human BAG3(P209L) in mice causes restrictive cardiomyopathy. Nat. Commun. 2021, 12, 3575. [Google Scholar] [CrossRef]
  223. Fang, X.; Bogomolovas, J.; Zhou, P.S.; Mu, Y.; Ma, X.; Chen, Z.; Zhang, L.; Zhu, M.; Veevers, J.; Ouyang, K.; et al. P209L mutation in Bag3 does not cause cardiomyopathy in mice. Am. J. Physiol. Heart Circ. Physiol. 2019, 316, H392–H399. [Google Scholar] [CrossRef]
  224. Kobuke, K.; Furukawa, Y.; Sugai, M.; Tanigaki, K.; Ohashi, N.; Matsumori, A.; Sasayama, S.; Honjo, T.; Tashiro, K. ESDN, a novel neuropilin-like membrane protein cloned from vascular cells with the longest secretory signal sequence among eukaryotes, is up-regulated after vascular injury. J. Biol. Chem. 2001, 276, 34105–34114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Schmoker, A.M.; Ebert, A.M.; Ballif, B.A. The DCBLD receptor family: Emerging signaling roles in development, homeostasis and disease. Biochem. J. 2019, 476, 931–950. [Google Scholar] [CrossRef] [PubMed]
  226. Sadeghi, M.M.; Esmailzadeh, L.; Zhang, J.; Guo, X.; Asadi, A.; Krassilnikova, S.; Fassaei, H.R.; Luo, G.; Al-Lamki, R.S.; Takahashi, T.; et al. ESDN is a marker of vascular remodeling and regulator of cell proliferation in graft arteriosclerosis. Am. J. Transplant. 2007, 7, 2098–2105. [Google Scholar] [CrossRef] [Green Version]
  227. Nie, L.; Guo, X.; Esmailzadeh, L.; Zhang, J.; Asadi, A.; Collinge, M.; Li, X.; Kim, J.D.; Woolls, M.; Jin, S.W.; et al. Transmembrane protein ESDN promotes endothelial VEGF signaling and regulates angiogenesis. J. Clin. Investig. 2013, 123, 5082–5097. [Google Scholar] [CrossRef]
  228. Xie, P.; Yuan, F.Q.; Huang, M.S.; Zhang, W.; Zhou, H.H.; Li, X.; Liu, Z.Q. DCBLD2 affects the development of colorectal cancer via emt and angiogenesis and modulates 5-FU drug resistance. Front. Cell Dev. Biol. 2021, 9, 669285. [Google Scholar] [CrossRef] [PubMed]
  229. Olivotto, I.; Oreziak, A.; Barriales-Villa, R.; Abraham, T.P.; Masri, A.; Garcia-Pavia, P.; Saberi, S.; Lakdawala, N.K.; Wheeler, M.T.; Owens, A.; et al. Mavacamten for treatment of symptomatic obstructive hypertrophic cardiomyopathy (EXPLORER-HCM): A randomised, double-blind, placebo-controlled, phase 3 trial. Lancet 2020, 396, 759–769. [Google Scholar] [CrossRef]
  230. Teerlink, J.R.; Diaz, R.; Felker, G.M.; McMurray, J.J.V.; Metra, M.; Solomon, S.D.; Adams, K.F.; Anand, I.; Arias-Mendoza, A.; Biering-Sorensen, T.; et al. Cardiac myosin activation with omecamtiv mecarbil in systolic heart failure. N. Engl. J. Med. 2021, 384, 105–116. [Google Scholar] [CrossRef]
  231. Jinek, M.; Chylinski, K.; Fonfara, I.; Hauer, M.; Doudna, J.A.; Charpentier, E. A programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity. Science 2012, 337, 816–821. [Google Scholar] [CrossRef]
  232. Ozcan, A.; Krajeski, R.; Ioannidi, E.; Lee, B.; Gardner, A.; Makarova, K.S.; Koonin, E.V.; Abudayyeh, O.O.; Gootenberg, J.S. Programmable RNA targeting with the single-protein CRISPR effector Cas7-11. Nature 2021, 597, 720–725. [Google Scholar] [CrossRef]
  233. Jungmann, A.; Leuchs, B.; Rommelaere, J.; Katus, H.A.; Muller, O.J. Protocol for efficient generation and characterization of adeno-associated viral vectors. Hum. Gene Ther. Methods 2017, 28, 235–246. [Google Scholar] [CrossRef]
  234. Weinmann, J.; Weis, S.; Sippel, J.; Tulalamba, W.; Remes, A.; El Andari, J.; Herrmann, A.K.; Pham, Q.H.; Borowski, C.; Hille, S.; et al. Identification of a myotropic AAV by massively parallel in vivo evaluation of barcoded capsid variants. Nat. Commun. 2020, 11, 5432. [Google Scholar] [CrossRef] [PubMed]
  235. Fomin, A.; Gartner, A.; Cyganek, L.; Tiburcy, M.; Tuleta, I.; Wellers, L.; Folsche, L.; Hobbach, A.J.; von Frieling-Salewsky, M.; Unger, A.; et al. Truncated titin proteins and titin haploinsufficiency are targets for functional recovery in human cardiomyopathy due to TTN mutations. Sci. Transl. Med. 2021, 13, eabd3079. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (A) Apical four chamber view during systole of an echocardiogram (B) and four chamber view of cardiac magnetic resonance image of a 50-year-old patient carrying a pathogenic FLNC mutation. Note the enlarged atria, normal ventricular sizes and wall thicknesses. RA = right atrium; RV = right ventricle; LA = left atrium; LV = left ventricle.
Figure 1. (A) Apical four chamber view during systole of an echocardiogram (B) and four chamber view of cardiac magnetic resonance image of a 50-year-old patient carrying a pathogenic FLNC mutation. Note the enlarged atria, normal ventricular sizes and wall thicknesses. RA = right atrium; RV = right ventricle; LA = left atrium; LV = left ventricle.
Jcm 11 02094 g001
Figure 2. Overview of RCM genes. (A) Genes associated with restrictive cardiomyopathy (RCM) according to the year of discovery. Different subcellular localizations are color-coded (red = sarcomere; green = cytoskeleton; yellow = Z-disc and blue = others). (B) Chromosomal location of RCM-associated genes. Schematic idiograms were licensed from shutterstock.de.
Figure 2. Overview of RCM genes. (A) Genes associated with restrictive cardiomyopathy (RCM) according to the year of discovery. Different subcellular localizations are color-coded (red = sarcomere; green = cytoskeleton; yellow = Z-disc and blue = others). (B) Chromosomal location of RCM-associated genes. Schematic idiograms were licensed from shutterstock.de.
Jcm 11 02094 g002
Figure 3. Venn diagram showing the genetic overlap of restrictive cardiomyopathy (RCM) with other cardiomyopathies. ACM = arrhythmogenic cardiomyopathy; DCM = dilated cardiomyopathy; HCM = hypertrophic cardiomyopathy; LVNC = left ventricular non-compaction cardiomyopathy; and MFM = myofibrillar myopathy. Gene names according to the HUGO Gene Nomenclature Committee, HGNC (https://www.genenames.org/ (accessed on 13 March 2022)). Sub-images of the DCM or HCM heart were licensed from shutterstock.de.
Figure 3. Venn diagram showing the genetic overlap of restrictive cardiomyopathy (RCM) with other cardiomyopathies. ACM = arrhythmogenic cardiomyopathy; DCM = dilated cardiomyopathy; HCM = hypertrophic cardiomyopathy; LVNC = left ventricular non-compaction cardiomyopathy; and MFM = myofibrillar myopathy. Gene names according to the HUGO Gene Nomenclature Committee, HGNC (https://www.genenames.org/ (accessed on 13 March 2022)). Sub-images of the DCM or HCM heart were licensed from shutterstock.de.
Jcm 11 02094 g003
Figure 5. Schematic overview of RCM associated DES mutations. (A) Schematic overview about the DES gene consisting of nine exons (NM_001927.4). Three splice site mutations have been identified in RCM patients at the donor splice site of exon 3. (B) Schematic domain organization of desmin and the localization of the known RCM-associated DES missense mutations.
Figure 5. Schematic overview of RCM associated DES mutations. (A) Schematic overview about the DES gene consisting of nine exons (NM_001927.4). Three splice site mutations have been identified in RCM patients at the donor splice site of exon 3. (B) Schematic domain organization of desmin and the localization of the known RCM-associated DES missense mutations.
Jcm 11 02094 g005
Figure 7. Schematic overview about the domain organization of filamin-C and the localization of the known RCM-associated FLNC missense mutations.
Figure 7. Schematic overview about the domain organization of filamin-C and the localization of the known RCM-associated FLNC missense mutations.
Jcm 11 02094 g007
Figure 8. Molecular structure of the αB-crystallin domain determined by nuclear magnetic resonance (NMR) spectroscopy (https://www.rcsb.org/structure/2KLR) (accessed on 13 March 2022) [206]. Two ion bridges are formed between aspartate p.D109 (blue) and arginine p.R120 (yellow) mediating its dimerization. Of note, the mutation CRYAB-p.D109G is associated with RCM in combination with skeletal myopathy [40] and -p.R120G causes MFM in combination with HCM and cataract [19,207].
Figure 8. Molecular structure of the αB-crystallin domain determined by nuclear magnetic resonance (NMR) spectroscopy (https://www.rcsb.org/structure/2KLR) (accessed on 13 March 2022) [206]. Two ion bridges are formed between aspartate p.D109 (blue) and arginine p.R120 (yellow) mediating its dimerization. Of note, the mutation CRYAB-p.D109G is associated with RCM in combination with skeletal myopathy [40] and -p.R120G causes MFM in combination with HCM and cataract [19,207].
Jcm 11 02094 g008
Table 5. Overview about the known RCM-associated FLNC mutations.
Table 5. Overview about the known RCM-associated FLNC mutations.
MutationClinical FeaturesFamily HistoryMAF 1CommentsReferences
p.A1183LRCM and congenital myopathyone patient- [176]
p.A1186VRCM and congenital myopathythree unrelated index patients-de novo[176]
RCMone patient-de novo,
early onset
[182]
p.S1624LRCMfour affected family members0.00003 [37]
p.C1639YRCMone patient-de novo,
early onset
[182]
p.G2011RRCMone patient-iPSC model[180]
p.G2151SRCMtwo patients-in addition PTPN11-p.Q510R[183]
p.I2160FRCMthree affected family members- [37]
p.V2264MRCM, SMone patient-iPSC model[179]
p.V2297MRCM, AFfive affected family members0.000004 [181]
p.P2298LRCMeight patients (four genotyped)- [184]
p.P2301LRCM, AF, muscular weaknessone patient-de novo[183]
p.Y2563CRCMtwo monozygotic twins-de novo[184]
1 MAF = Minor allele frequency according to Genome Aggregation Database (January 2022), https://gnomad.broadinstitute.org/(accessed on 13 March 2022). AF = atrial fibrillation, RCM = restrictive cardiomyopathy, SM = skeletal myopathy.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Brodehl, A.; Gerull, B. Genetic Insights into Primary Restrictive Cardiomyopathy. J. Clin. Med. 2022, 11, 2094. https://doi.org/10.3390/jcm11082094

AMA Style

Brodehl A, Gerull B. Genetic Insights into Primary Restrictive Cardiomyopathy. Journal of Clinical Medicine. 2022; 11(8):2094. https://doi.org/10.3390/jcm11082094

Chicago/Turabian Style

Brodehl, Andreas, and Brenda Gerull. 2022. "Genetic Insights into Primary Restrictive Cardiomyopathy" Journal of Clinical Medicine 11, no. 8: 2094. https://doi.org/10.3390/jcm11082094

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop