Skip to content
Publicly Available Published by De Gruyter September 2, 2014

Titin: central player of hypertrophic signaling and sarcomeric protein quality control

  • Sebastian Kötter

    After finishing his Biology studies at Münster University in 2007, Sebastian Kötter joined the Titin lab of Wolfgang Linke and obtained his PhD on the sarcomeric organization and phosphorylation of the titin filament, and its association to small heat shock proteins. In 2011 he joined the lab of Martina Krüger as a post doc and continued his work on titin modification and regulation of titin turnover.

    , Christian Andresen

    Christian Andresen he received his diploma degree in biology at the University of Münster in 2004. He continued his scientific work in the lab of Wolfgang Linke and got his PhD in 2008. His research focused on studying the sarcomere structure and stability and on the investigation of titin-protein interactions.

    and Martina Krüger

    After graduating as a biologist in 1999, Martina Krüger obtained her PhD in 2004 at the Physiology Department of the University of Cologne. In 2005 she joined the Titin lab of Wolfgang Linke as a postdoc to study titin-based myofilament function in health and disease. In 2011 she became a full professor at Düsseldorf University. Her current research focuses on posttranslational modification of titin and the hormonal regulation of titin-based myofilament properties.

    EMAIL logo
From the journal Biological Chemistry

Abstract

The giant sarcomeric protein titin has multiple important functions in striated muscle cells. Due to its gigantic size, its central position in the sarcomere and its elastic I-band domains, titin is a scaffold protein that is important for sarcomere assembly, and serves as a molecular spring that defines myofilament distensibility. This review focuses on the emerging role of titin in mechanosensing and hypertrophic signaling, and further highlights recent evidence that links titin to sarcomeric protein turnover.

Introduction

The sarcomere is the smallest contractile unit of striated muscle cells and is mainly composed of three filament systems: the myosin-based thick filament; the actin-based thin filament, supplemented with the regulatory protein tropomyosin and the troponin complex; and the titin filament. The giant protein titin is encoded by a single titin gene and has a potential size of 4.2 MDa, allowing one titin molecule to span an entire half-sarcomere from the Z-disc to the M-line. Differential splicing of the titin gene is the basis for numerous species- and muscle-specific titin isoforms with molecular weights ranging from 3.0 MDa to 3.7 MDa. Most of the alternative splicing events occur in the I-band region of titin, but to a smaller degree they also affect the Z-disc and the M-band portion of the molecule (Krüger and Linke, 2011). The molecular mechanisms that control the complex splicing events of the titin mRNA are still not completely understood, but recent studies have identified the splicing factor RBM20 as an important player in this setting (Guo et al., 2012; Li et al., 2013).

The mammalian heart expresses two main isoform types of titin: the longer and more compliant N2BA isoforms (3.2–3.7 MDa) and the shorter and stiffer N2B isoform (3.0 MDa). During cardiac development the heart expresses a compliant fetal N2BA isoform that is replaced after birth by the smaller N2BA and the N2B isoform (Opitz et al., 2004). Depending on the species, this isoform shift is more or less pronounced (Krüger et al., 2006); in adult rat hearts the N2B isoform predominates, with approximately 90% of total titin; whereas in healthy human hearts the relative expression ratio is about 35% N2BA:65% N2B (Neagoe et al., 2002). Skeletal muscles express a third isoform type called N2A titin (3.3–3.7 MDa) with many muscle-specific splice variants (Freiburg et al., 2000; Neagoe et al., 2003; Prado et al., 2005).

Structure and function of titin

The structural composition of titin is mainly characterized by the sequential arrangement of immunoglobulin-like domains (Ig-domains), fibronectin-type-3 domains and several so-called unique sequences (Bang et al., 2001). The NH2-terminal end of titin is anchored in the sarcomeric Z-disc via nebulin or the cardiac isoform nebulette (Witt et al., 2006), α-actinin 2 (Labeit et al., 2006) and telethonin (Granzier and Labeit, 2004; Miller et al., 2004; Lange et al., 2006). In the I-band part titin is composed of a series of proximal Ig-domains, the cardiac specific N2B-domain (including the N2-B unique sequence, N2-Bus), the middle Ig-domain region (not in the N2B isoform), the PEVK domain [predominant amino acids: proline (P), glutamic acid (E), valine (V) and lysine (K)] and the distal Ig-domains. This I-band part can be sequentially extended during sarcomere stretch and represents the main elastic segment of titin (Linke et al., 1996, 1999; Trombitas et al., 1998; Li et al., 2002). With a size of about 2 MDa, the A-band portion of titin is the largest part of the molecule (Bang et al., 2001) and it is tightly associated with myosin and myosin-binding protein C (Tskhovrebova and Trinick, 2004; Lange et al., 2006). The M-band portion of titin is characterized by several inserted sequences and the titin–kinase-domain in the M-band periphery (Figure 1) (Bang et al., 2001; Gautel, 2011).

Figure 1 Titin-mediated mechanosensing and hypertrophic signaling.Various titin–protein interactions at the Z-disc, in the I-band, and in the M-band link titin to hypertrophic signaling pathways. CaM, Ca2+-calmodulin; ERK 1/2, extracellular signal-regulated kinase; FHL1/2, four-and-a-half LIM domain protein 1/2; Fn-like domain, fibronectin-like domain; Ig-like domain, immunoglobulin-like domain; MEK 1/2, mitogen-activated protein kinase kinase 1; MLP, muscle LIM protein; MuRF 1/2, muscle ring finger protein 1/2; Nbr1, neighbor-of-BRCA1-gene-1; NFAT, nuclear factor of activated T-cell; p62, nucleoporin p62; Raf, rapidly accelerated fibrosarcoma protein; SQSTM1, sequestosome 1; telethonin, telethonin (titin-cap).
Figure 1

Titin-mediated mechanosensing and hypertrophic signaling.

Various titin–protein interactions at the Z-disc, in the I-band, and in the M-band link titin to hypertrophic signaling pathways. CaM, Ca2+-calmodulin; ERK 1/2, extracellular signal-regulated kinase; FHL1/2, four-and-a-half LIM domain protein 1/2; Fn-like domain, fibronectin-like domain; Ig-like domain, immunoglobulin-like domain; MEK 1/2, mitogen-activated protein kinase kinase 1; MLP, muscle LIM protein; MuRF 1/2, muscle ring finger protein 1/2; Nbr1, neighbor-of-BRCA1-gene-1; NFAT, nuclear factor of activated T-cell; p62, nucleoporin p62; Raf, rapidly accelerated fibrosarcoma protein; SQSTM1, sequestosome 1; telethonin, telethonin (titin-cap).

Titin has long been recognized as a scaffolding protein that serves as a ‘blue print’ for sarcomerogenesis and assists in the process of myofibrillar assembly (Ehler and Gautel, 2008; Tskhovrebova and Trinick, 2010). The Ig-domains and fibronectin-type-3 domains in the A-band are arranged in seven- and 11-domain super repeats, which are repeated six and 11 times, respectively. A detailed overview of titin domain structure in the different parts of the sarcomere is provided in Krüger and Linke (2011).

Another important function of titin is provided by the possibility of reversibly extending the elastic I-band domains upon mechanical stretching, thus allowing titin to act as a molecular spring and thereby defining the passive properties of the myofilaments. In cardiac myocytes the titin-based passive myofilament stiffness is mainly determined by the expression ratio of the N2BA and N2B titin isoforms (Krüger and Linke, 2011). In end-stage heart failure, the ratio of N2BA:N2B is often increased and results in a reduced passive stiffness (Neagoe et al., 2002; Makarenko et al., 2004; Nagueh et al., 2004). It has, however, also been shown that in some end-stage dilated and hypertrophic cardiomyopathy patients the titin isoform ratio remains unchanged and passive stiffness increases due to alterations in the titin phosphorylation status (Kötter et al., 2013).

Posttranslational modification of titin modulates titin stiffness more dynamically, e.g., by phosphorylation of the elastic I-band regions N2-B and PEVK. To date, phosphorylation sites have been identified for cAMP-dependent protein kinase (Yamasaki et al., 2002; Krüger et al., 2009; Kötter et al., 2013), cGMP-dependent protein kinase (Krüger et al., 2009; Kötter et al., 2013), Ca2+-dependent protein kinase C α (Hidalgo et al., 2009), extracellular signal regulated kinase 1/2 (Erk1/2) (Raskin et al., 2012) and Ca2+/calmodulin-dependent protein kinase II delta (CaMKIIδ) (Hamdani et al., 2013b; Hidalgo et al., 2013). Phosphorylation of the cardiac-specific N2-Bus by cAMP- and cGMP-dependent protein kinase and CaMKIIδ increases the persistence length of this region, and thereby decreases the passive stiffness (Krüger et al., 2009; Hamdani et al., 2013b). In contrast, phosphorylation of the PEVK domain by Ca2+-dependent protein kinase C α decreases the persistence length of the PEVK region and causes an increase in titin-based passive tension (Hidalgo et al., 2009). Why the addition of phosphates results in an increase of the persistence length in one titin region whereas it causes a decrease in another is not fully understood. One possible explanation is the different amino acid composition of the domains. The introduction of a negatively-charged phosphate group into an already negatively-charged environment, such as the N2-Bus, can have an opposite impact to introducing it into a positively-charged sequence, such as the PEVK region. This could lead to intramolecular electrostatic repulsion, thereby altering the persistence length of the domains (Kötter et al., 2013). Chronic changes in the phosphorylation status of titin without concomitant changes in titin isoform expression have been observed in different studies analyzing biopsies from heart failure patients and have resulted in severe alterations of titin-based myofilament stiffness (Kötter et al., 2013; Hamdani et al., 2013a). The role of titin phosphorylation in modulating diastolic function has recently been reviewed in more detail (Linke and Hamdani, 2014). Recent evidence from in vitro studies showed posttranslational modification of titin by S-glutathionylation, and a subsequent reduction in titin-based myofilament stiffness (Avner et al., 2012; Alegre-Cebollada et al., 2014). This modification raises the possibility that myofilament stiffness could be modulated by S-glutathionylation under acidic conditions, such as cardiac ischemia. The analysis of titin modification is an emerging field, however, and will doubtlessly result in the identification of many more modifications of titin and titin function in the near future.

Apart from its importance for sarcomeric stiffness, titin plays an important role as a mediator of myocyte signal transduction. The following section will place special emphasis on titin’s putative involvement in hypertrophic signaling and its role in the complex regulatory mechanisms of protein quality control.

A mediator of mechanosensing and hypertrophic signaling

Sensing of mechanical stress and regulation of specific stretch responses is an important and highly conserved function in both cardiac and skeletal muscle. The adaptation of muscle function to increasing mechanical demands involves the initiation of a hypertrophic gene program to increase the number of sarcomeres, and often requires re-activation of fetal gene expression (Hunter and Chien, 1999). The existence of a specific signaling pathway that couples mechanical stress and muscle gene expression seems very likely; however, the molecular mechanism of this process is still unresolved.

Due to its large size and the longitudinal position within the sarcomere, titin is a very potent candidate sensor of mechanical load. Titin–protein interaction has been reported for more than 20 different proteins and is localized to three major sites of the molecule: the Z-disc region, the elastic I-band domains N2-B and N2-A, and the M-band proximity including the titin kinase domain (for a more detailed review of titin–protein interactions, see Linke and Krüger, 2010). Some of the identified binding partners of titin have been linked to hypertrophic signaling.

The Z-disc part of titin is mainly composed of the so-called Z-repeats and the Ig-domains Z1 and Z2, which form the very NH2-terminal end of titin that protrudes from the Z-disc region. Telethonin connects the NH2-terminal parts of two titin molecules from one sarcomere in a palindromic manner (Zou et al., 2006), an interaction that has been reported to be essential for proper sarcomere integrity (Gregorio et al., 1998). Cardiac telethonin is a substrate for atypical members of the Ca2+/calmodulin-dependent kinase family, including titin kinase, CaMKII, and protein kinase D. Recent evidence suggests that constitutive bis-phosphorylation is critical for normal telethonin function that further involves maintenance of transverse tubule organization and intracellular Ca2+ transients (Candasamy et al., 2014). Missense mutations in telethonin have been associated with limb girdle muscular dystrophy 2G in skeletal muscle (Moreira et al., 2000) and with hypertrophic and dilated cardiomyopathies (Hayashi et al., 2004). Interestingly, some of the reported missense mutations of telethonin may interfere with the phosphorylation of telethonin and thereby disturb its proper function (Candasamy et al., 2014). An important binding partner of telethonin is the muscle LIM protein MLP, a zinc-finger protein of the LIM-only protein family, which has been shown in MLP-deficient mice to play a crucial role in myogenic differentiation (Arber et al., 1997). In conjunction with the Z-disc component actinin, it has been proposed that MLP, titin and telethonin form a mechanical stretch sensing complex (Knöll et al., 2002, 2011) (Figure 1). MLP contains a nuclear localization signal, allowing cytoplasmic as well as nuclear localization of the protein. Increased nucleo-cytoplasmic shuttling of MLP has been observed in response to biomechanical stress and has been associated with the initiation of hypertrophic remodeling in stressed myocytes (Boateng et al., 2009). As MLP lacks a DNA binding site, however, the precise function of nuclear MLP remains to be elucidated (for more details on MLP, see Buyandelger et al., 2011). The importance of MLP has been stressed by the identification of several missense mutations associated with dilated cardiomyopathy, possibly caused by loss or disturbance of the MLP/telethonin/titin sensor complex (Knöll et al., 2010).

The MLP/telethonin/titin sensor complex also provides a direct link to hypertrophic signaling, which involves activation of the calcineurin/nuclear factor of activated T-cells (NFAT) signaling cascade. Telethonin has been shown to interact with calsarcin, a protein that tethers calcineurin to the Z-disc (Figure 1) (Frey et al., 2000). Calcineurin is a calcium/calmodulin-dependent serine/threonine phosphatase, which plays an important role in striated muscle signal transduction. The main function of calcineurin is the dephosphorylation of NFAT, thereby inducing NFAT translocation to the nucleus and activation of target genes (for a review, see Olson and Williams, 2000). Constitutive activation of calcineurin in the hearts of transgenic mice induces severe hypertrophy and leads to heart failure and sudden death. Evidence suggests that MLP plays a crucial role in the activation of calcineurin when associated with the sarcomeric Z-disc (Heineke et al., 2005).

The main site of protein–protein interaction in the I-band is the elastic cardiac-specific N2-B region, which consists of several Ig-domains and a so-called unique sequence (N2-Bus). The N2-Bus has been shown to bind the two isoforms of the four-and-a-half LIM domain protein FHL-1 (Sheikh et al., 2008) and FHL-2 (Lange et al., 2002). Both isoforms are highly expressed in response to biomechanical stress (Lim et al., 2001; Sheikh et al., 2008), can shuttle to the nucleus and act as transcriptional co-activators (Scholl et al., 2000). FHL-2 is expressed early during cardiogenesis and remains highly expressed in adult tissue; however, its function in the heart is not entirely understood. Recent data from a FHL-2 knockout mouse model and from cellular overexpression studies suggest that FHL-2 is an endogenous suppressor of calcineurin and thereby represses pathological cardiac growth (Hojayev et al., 2012). Apart from its I-band localization, FHL-2 has been shown to associate with the myofilament at the M-band and the Z-disc. In contrast, FHL1 association to titin may provide a link to the mitogen-activated protein kinase (MAPK) signaling cascade (Raf/Mek1/2/Erk2). Under non-stimulating conditions MEK1/2 acts as a scaffold protein for ERK1/2, and anchors ERK in the cytoplasm (Tanoue et al., 2000; Chuderland and Seger, 2005). Upon activation, MEK1/2 phosphorylates ERK1/2 and induces a conformational change that leads to its dissociation and translocation to the nucleus, where ERK2 is responsible for the activation of transcription factors such as c-Myc, c-Fos, and cAMP response element-binding (Mebratu and Tesfaigzi, 2009). ERK2 has further been shown to phosphorylate titin’s N2-Bus sequence in vitro (Raskin et al., 2012) and possibly modulate myofilament stiffness. Interestingly, knockdown of FHL1 in mice resulted in increased myofibrillar compliance and reduced hypertrophic signaling (Sheikh et al., 2008). The N2-B/FHL-1/MAPK complex may therefore represent another potent biomechanical stress sensor in cardiomyocytes (Figure 1).

The third hotspot for titin-mediated hypertrophic signaling is the M-band region of the molecule, especially the titin kinase (TK) domain located in the M-band periphery. To date, the only substrate of the TK domain that has been identified is telethonin (Mayans et al., 1998; Weinert et al., 2006). Structurally, the titin kinase has some similarities with other kinases of the myosin light-chain kinase family, and shows a strong auto-inhibition (Gautel, 2011). Unlike other classical myosin light-chain kinases, however, the modulation of TK activity by Ca2+-calmodulin is only very weak (Mayans et al., 1998). Instead, it has been proposed from force-probe molecular dynamics simulations that activating changes to the conformational state of the TK could be induced biomechanically (Gräter et al., 2005). The TK therefore could be another biomechanical stress sensor and mediator of hypertrophic signaling.

Activated TK has been shown to interact with the ubiquitin-associated zinc-finger protein neighbor-of-BRCA1-gene-1 (Nbr1), which forms a signaling complex with p62/SQSTM1 and the muscle-specific ubiquitin E3 ligases MuRF1, MuRF2, and MuRF3 (Lange et al., 2005). It has been suggested that MuRFs have act as transcriptional repressors in mechanically inactive cardiac and skeletal muscle cells (Gautel, 2011). In the absence of mechanical activity, MURFs shuttle to the nucleus, where MuRF2 has the ability to down-regulate the levels of nuclear serum response factor and thereby suppress serum response factor-dependent muscle gene expression (Lange et al., 2005). By interactions with several members of the MAPK signaling pathway, including p38, P62/SQSTM1 is likely involved in myocyte enhancer factor-2- and MyoD-controlled myogenic differentiation and hypertrophic growth (reviewed in Gautel, 2011). Interestingly, a human mutation in the titin protein kinase domain that disrupts the formation of this signaling complex has been shown to cause severe hereditary muscle disease (Lange et al., 2005).

A central part of sarcomeric protein quality control

Sarcomeres are highly dynamic structures that require permanent synthesis and turnover of their constituent parts (McKenna et al., 1985; Sanger et al., 1986; Mittal et al., 1987). There are three main turnover processes in the cell: the ubiquitin–proteasome system (UPS), autophagy and the calpains. The degradation of muscle proteins by the UPS plays an important role in the healthy heart and is often disturbed in patients with cardiac diseases, such as heart failure (Tsukamoto et al., 2006; Wang and Robbins, 2006; Predmore et al., 2010; Wohlschlaeger et al., 2010; Watkins et al., 2011), cardiomyopathies (Liu et al., 2006; Carrier et al., 2010, Predmore et al., 2010; Schlossarek and Carrier, 2011), hypertrophy (Hein et al., 2003; Depre et al., 2006; Meiners et al., 2008), atrophy (Attaix et al., 2005; Razeghi et al., 2006), ischemia-reperfusion (Powell and Divald, 2010; Li et al., 2011) and atherosclerosis (Herrmann et al., 2004). Despite this, autophagy is increasingly recognized as an important regulator of protein degradation and cellular signal transduction.

The following paragraphs will highlight the emerging evidence suggesting that the giant molecule titin represents an important signaling node in the processes of sarcomeric protein quality control, and at the same time could be its biggest challenge.

In the past several studies have directly or indirectly linked titin to important players of the protein quality control machinery and the ubiquitin proteasome system. In the Z-disc a potent connection to the proteasomal system is provided by the interaction of the ubiquitin–ligase Mice-double-minute 2 (Mdm2) with the titin-capping protein telethonin. The titin M-band domains A168–170 interact with MuRF-1 and MuRF-2, E3-ligases that are involved in the proteasomal degradation of several muscle proteins including troponins, telethonin and nebulin (Pizon et al., 2002; Centner et al., 2003; Gregorio et al., 2005; Witt et al., 2005) (Figure 2). Through MuRF-1, titin is also linked to the ubiquitin binding protein 9 (McElhinny et al., 2002).

Figure 2 Sarcomeric protein quality control and chaperone-mediated titin protection.Titin has been demonstrated to interact with several members of the protein quality control machinery. Interaction with E3 ubiquitin ligases provides a potent link to the ubiquitin proteasome system, whereas the titin-associated Nbr1/p62/SQSTM1 complex may promote the autophagosomal degradation of ubiquitinated proteins. Inset: chaperone-mediated protection of I-band titin.HSP27, heat shock protein 27; HSP90, heat shock protein 90; LC3, microtubule-associated protein 1 light chain 3; Mdm2, Mice-double-minute 2; Nbr1, neighbor-of-BRCA1-gene-1; p62, nucleoporin p62; Smyd2, SET and MYND domain-containing protein 2; SQSTM1, sequestosome 1; telethonin, telethonin (titin-cap).
Figure 2

Sarcomeric protein quality control and chaperone-mediated titin protection.

Titin has been demonstrated to interact with several members of the protein quality control machinery. Interaction with E3 ubiquitin ligases provides a potent link to the ubiquitin proteasome system, whereas the titin-associated Nbr1/p62/SQSTM1 complex may promote the autophagosomal degradation of ubiquitinated proteins. Inset: chaperone-mediated protection of I-band titin.

HSP27, heat shock protein 27; HSP90, heat shock protein 90; LC3, microtubule-associated protein 1 light chain 3; Mdm2, Mice-double-minute 2; Nbr1, neighbor-of-BRCA1-gene-1; p62, nucleoporin p62; Smyd2, SET and MYND domain-containing protein 2; SQSTM1, sequestosome 1; telethonin, telethonin (titin-cap).

Another important link between titin and protein turnover processes is mediated by the interaction of the titin-kinase domain with the Nbr1/p62/SQSTM1 complex and the subsequent recruitment of MuRF-2 to the sarcomere (Figure 2) (Lange et al., 2005). By targeting ligands to polyubiquitin chains, P62/SQSTM1 can induce the assembly of larger signalosomes and promote their proteasomal degradation (Seibenhener et al., 2007). Via interaction with the autophagosomal membrane anchor LC3, p62 and Nbr1 have been shown to target polyubiquitinated proteins to the autophagic protein turnover machinery (Pankiv et al., 2007; Waters et al., 2009). These data demonstrate that, apart from its role in hypertrophic signal transduction, p62/SQSTM1/Nbr1 is an important signaling complex in the crosstalk between the UPS and autophagosomal processes.

A direct contribution of titin to protein quality control mechanisms is provided by the association of the proteases calpain-1 to the proximal Ig-region of titin, and calpain-3 (skeletal muscle-specific isoform) to the N2A-domain and the M-band region of titin (Raynaud et al., 2005; Coulis et al., 2008; Hayashi et al., 2008). It has been suggested that the binding to titin keeps calpain-3 in an autoinhibited state and thereby regulates its proteolytic activity (Hayashi et al., 2008). Moreover, calpains, and particularly calpain-1, seem to be required for ubiquitin ligases to mark sarcomeric proteins for degradation by the proteasome (Fareed et al., 2006).

Much faster than expected from its giant size, titin can be rapidly exchanged in the sarcomere within approximately 14 h, which is probably enabled by a cytoplasmic pool of titin (da Silva-Lopes et al., 2011). How these processes allow an accurate integration of titin in the working sarcomere is still entirely unknown. It is also still unresolved whether titin degradation is mainly performed by the proteasome or the autophagosomal system or a combination of both. As titin is firmly integrated in the Z-disc, the A-band and M-line structure, it seems likely that the giant molecule is subjected to some kind of pre-digestion that allows subsequent proteasomal degradation. This has been demonstrated earlier in models of skeletal muscle cachexia, sepsis and muscle wasting. The pre-digestion has been performed by calpains (Williams et al., 1999; Hasselgren and Fischer, 2001). More recently, titin has been shown to be degraded by the matrix-metalloprotease (MMP) MMP-2 in vitro and in vivo, e.g., after cardiac ischemia/reperfusion injury (Ali et al., 2010). Although usually located in the extracellular matrix, the intracellular and particularly the sarcomeric localization of MMPs have previously been convincingly demonstrated. Under conditions of oxidative stress, MMP-2 is responsible for the degradation of several sarcomeric proteins like troponin, myosin-light chain and α-actinin (Wang et al., 2002; Gao et al., 2003; Sawicki et al., 2005; Sung et al., 2007).

Chaperone-mediated protection of titin

Protein quality control not only involves the degradation and recycling of proteins but also their stabilization and protection from damage. Titin has been shown to associate with important components of the protective protein quality control machinery: the heat shock proteins (HSPs). HSPs play a major role in protein homeostasis by protecting unfolded proteins from aggregation and keeping them in a state that allows refolding or degradation (Vos et al., 2008). In the following section we will highlight some recent findings demonstrating the association of titin with HSP90 and small HSPs (sHSPs).

In skeletal muscle the elastic N2A-domain in the I-band region of titin associates with the methyltransferase SET and MYND domain containing protein 2 (Smyd2), which in turn recruits HSP90 to this region and monomethylates it on lysine residue K616 (Donlin et al., 2012). This Smyd2-mediated monomethylation is crucial for the formation of a complex consisting of titin-N2A, Smyd2 and HSP90, which protects I-band titin against degradation. Deletion of Smyd2 in a zebrafish model resulted in severe I-band damage and loss of muscle function in skeletal as well as cardiac muscle (Donlin et al., 2012, Voelkel et al., 2013). In the mammalian system, however, knockdown of Smyd2 did not display any visible phenotype in skeletal or heart muscle (Diehl et al., 2010). A possible explanation for this discrepancy could be a functional compensation of Smyd2-KO by other members of the Smyd family, e.g., Smyd1, 3, 4 and 5 (Donlin et al., 2012). Nonetheless, these findings indicate that the complex consisting of titin-N2A/Smyd2/HSP90 could be important for sarcomere stabilization.

Recent studies have focused on the role of sHSPs in protecting the giant molecule titin from degradation. SHSPs are constitutively expressed in various tissues and their expression can be additionally up-regulated under stress situations (Lanneau et al., 2010). All members of the sHSP family exist in an oligomeric state that is usually induced by self-association of sHSP dimers (Mchaourab et al., 1997; Vos et al., 2008). Introducing a negative charge into the hydrophobic N-terminus of sHSPs, e.g., by phosphorylation, leads to dissociation into smaller oligomers and alters the chaperone activity (Ahmad et al., 2008; Vos et al., 2008). The main function of sHSPs is to bind partially unfolded proteins and hold them in a folding-prone state (Horwitz, 2003). If refolding fails the bound protein becomes ubiquitinated and degraded, predominantly by the proteasome (Vos et al., 2008).

Several sHSPs have been shown to translocate to the sarcomeric Z-disc and I-band under cellular stress conditions (Barbato et al., 1996; Lutsch et al., 1997; van de Klundert et al., 1998; Golenhofen et al., 1999; Fischer et al., 2002; Paulsen et al., 2009). The best-studied mammalian members of the sHSPs are αB-crystallin and HSP27. Both are highly expressed in muscle tissue under standard conditions (Kato et al., 1991) and have been shown to associate with the titin filament (Bullard et al., 2004; Kötter et al., 2014). Interestingly, αB-crystallin and HSP27 specifically translocate to the elastic I-band region of titin under ischemic (acidic) conditions (Golenhofen et al., 2002; Bullard et al., 2004; Kötter et al., 2014) (Figure 2). Ig-domains from the I-band portion of titin differ largely concerning their intrinsic stability. In single molecule measurements, the proximal Ig-domains and the Ig-domains in the N2-A domain required much lower unfolding forces than the distal Ig-domains (Rief et al., 1998; Li et al., 2002; Watanabe et al., 2002; Li and Fernandez, 2003). Interestingly, the presence of αB-crystallin in such experiments elevated the unfolding forces of the titin Ig-domain constructs analyzed (Bullard et al., 2004; Zhu et al., 2009). A recent study demonstrated that under conditions of ischemic stress unfolding of the weaker proximal Ig-domains is easily induced by mechanical stretch, thereby inducing aggregation of titin filaments and a subsequent increase in passive myocyte stiffness. Importantly, the presence of the sHSPs αB-crystallin and HSP27 prevented the unfolding and aggregation of titin and may therefore represent an important mechanism to protect myocytes from abnormal passive stiffness under ischemic conditions (Kötter et al., 2014).

Conclusion

This review has briefly highlighted the recent data implicating a multifaceted role of titin in mediating hypertrophic signaling processes and sarcomeric protein turnover. Nevertheless, details of titin-associated downstream and upstream signaling pathways that maintain and modify sarcomeric structure and function under stress conditions are only just beginning to be elucidated. Future studies will certainly provide more mechanistic insight into the role of titin as a sarcomeric signaling node and will also shed light on the precise mechanisms that control and mediate integration, degradation, and possibly recycling of the titin molecule itself.


Corresponding author: Martina Krüger, Department of Cardiovascular Physiology, Heinrich Heine University Düsseldorf, D-40225 Düsseldorf, Germany, e-mail:

About the authors

Sebastian Kötter

After finishing his Biology studies at Münster University in 2007, Sebastian Kötter joined the Titin lab of Wolfgang Linke and obtained his PhD on the sarcomeric organization and phosphorylation of the titin filament, and its association to small heat shock proteins. In 2011 he joined the lab of Martina Krüger as a post doc and continued his work on titin modification and regulation of titin turnover.

Christian Andresen

Christian Andresen he received his diploma degree in biology at the University of Münster in 2004. He continued his scientific work in the lab of Wolfgang Linke and got his PhD in 2008. His research focused on studying the sarcomere structure and stability and on the investigation of titin-protein interactions.

Martina Krüger

After graduating as a biologist in 1999, Martina Krüger obtained her PhD in 2004 at the Physiology Department of the University of Cologne. In 2005 she joined the Titin lab of Wolfgang Linke as a postdoc to study titin-based myofilament function in health and disease. In 2011 she became a full professor at Düsseldorf University. Her current research focuses on posttranslational modification of titin and the hormonal regulation of titin-based myofilament properties.

Acknowledgments

Some of the work reviewed here was supported by grant KR 3409/5-1 from the German Research Foundation.

References

Ahmad, M.F., Raman, B., Ramakrishna, T., and Rao, C.M. (2008). Effect of phosphorylation on alpha B-crystallin: differences in stability, subunit exchange and chaperone activity of homo and mixed oligomers of alpha B-crystallin and its phosphorylation-mimicking mutant. J. Mol. Biol. 375, 1040–1051.10.1016/j.jmb.2007.11.019Search in Google Scholar PubMed

Alegre-Cebollada, J., Kosuri, P., Giganti, D., Eckels, E., Rivas-Pardo, J.A., Hamdani, N., Warren, C.M., Solaro, R.J., Linke, W.A., and Fernández, J.M. (2014). S-glutathionylation of cryptic cysteines enhances titin elasticity by blocking protein folding. Cell 156, 1235–1246.10.1016/j.cell.2014.01.056Search in Google Scholar PubMed PubMed Central

Ali, M.A., Cho, W.J., Hudson, B., Kassiri, Z., Granzier, H., and Schulz, R. (2010). Titin is a target of matrix metalloproteinase-2: implications in myocardial ischemia/reperfusion injury. Circulation 122, 2039–2047.10.1161/CIRCULATIONAHA.109.930222Search in Google Scholar PubMed PubMed Central

Arber, S., Hunter, J.J., Ross, J. Jr., Hongo, M., Sansig, G., Borg, J., Perriard, J.C., Chien, K.R., and Caroni, P. (1997). MLP-deficient mice exhibit a disruption of cardiac cytoarchitectural organization, dilated cardiomyopathy, and heart failure. Cell 88, 393–403.10.1016/S0092-8674(00)81878-4Search in Google Scholar PubMed

Attaix, D., Ventadour, S., Codran, A., Béchet, D., Taillandier, D., and Combaret, L. (2005). The ubiquitin-proteasome system and skeletal muscle wasting. Essays Biochem. 41, 173–186.10.1042/bse0410173Search in Google Scholar

Avner, B.S., Shioura, K.M., Scruggs, S.B., Grachoff, M., Geenen, D.L., Helseth, D.L. Jr., Farjah, M., Goldspink, P.H., and Solaro, R.J. (2012). Myocardial infarction in mice alters sarcomeric function via post-translational protein modification. Mol. Cell Biochem. 363, 203–215.10.1007/s11010-011-1172-zSearch in Google Scholar PubMed PubMed Central

Bang, M.L., Centner, T., Fornoff, F., Geach, A.J., Gotthardt, M., Mc- Nabb, M., Witt, C.C., Labeit, D., Gregorio, C.C., Granzier, H., et al. (2001). The complete gene sequence of titin, expression of an unusual approximately 700-kDa titin isoform, and its interaction with obscurin identify a novel Z-line to I-band linking system. Circ. Res. 89, 1065–1072.10.1161/hh2301.100981Search in Google Scholar PubMed

Barbato, R., Menabò, R., Dainese, P., Carafoli, E., Schiaffino, S., and Di Lisa, F. (1996). Binding of cytosolic proteins to myofibrils in ischemic rat hearts. Circ. Res. 78, 821–828.10.1161/01.RES.78.5.821Search in Google Scholar

Boateng, S.Y., Senyo, S.E., Qi, L., Goldspink, P.H., and Russel, B. (2009). Myocyte remodeling in response to hypertrophic stimuli requires nucelocytoplasmic shuttling of muscle LIM protein. J. Mol. Cell Cardiol. 47, 426–435.10.1016/j.yjmcc.2009.04.006Search in Google Scholar PubMed PubMed Central

Bullard, B., Ferguson, C., Minajeva, A., Leake, M.C., Gautel, M., Labeit, D., Ding, L., Labeit, S., Horwitz, J., Leonard, K.R., et al. (2004). Association of the chaperone αB-crystallin with titin in heart muscle. J. Biol. Chem. 279, 7917–7924.10.1074/jbc.M307473200Search in Google Scholar PubMed

Buyandelger, B., Ng, K.E., Miocic, S., Piotrowska, I., Gunkel, S., Ku, C.H., and Knöll, R. (2011). MLP (muscle LIM protein) as a stress sensor in the heart. Pflüger’s Arch. 462, 135–142.10.1007/s00424-011-0961-2Search in Google Scholar PubMed PubMed Central

Candasamy, A.J., Haworth, R.S., Cuello, F., Ibrahim, M., Aravamudhan, S., Krüger, M., Holt, M.R., Terracciano, C.M., Mayr, M., Gaute, l.M., et al. (2014). Phosphoregulation of the titin-cap protein telethonin in cardiac myocytes. J. Biol. Chem. 289, 1282–1293.10.1074/jbc.M113.479030Search in Google Scholar PubMed PubMed Central

Carrier, L., Schlossarek, S., Willis, M.S., and Eschenhagen, T. (2010). The ubiquitin-proteasome system and nonsense-mediated mRNA decay in hypertrophic cardiomyopathy. Cardiovasc. Res. 85, 330–338.10.1093/cvr/cvp247Search in Google Scholar PubMed PubMed Central

Centner, T., Yano, J., Kimura, E., McElhinny, A.S., Pelin, K., Witt, C.C., Bang, M.L., Trombitas, K., Granzier, H., Gregorio, C.C., et al. (2003). Identification of muscle specific ring finger proteins as potential regulators of the titin kinase domain. J. Mol. Biol. 306, 717–726.10.1006/jmbi.2001.4448Search in Google Scholar PubMed

Chuderland, D. and Seger, R. (2005). Protein-protein interactions in the regulation of the extracellular signal-regulated kinase. Mol. Biotechnol. 29, 57–74.10.1385/MB:29:1:57Search in Google Scholar PubMed

Coulis, G., Becila, S., Herrera-Mendez, C.H., Sentandreu, M.A., Raynaud, F., Richard, I., Benyamin, Y., and Ouali, A. (2008). Calpain 1 binding capacities of the N1-line region of titin are significantly enhanced by physiological concentrations of calcium. Biochemistry 47, 9174–9183.10.1021/bi800315vSearch in Google Scholar PubMed

da Silva-Lopes, K., Pietas, A., Radke, M.H., and Gotthardt, M. (2011). Titin visualization in real time reveals an unexpected level of mobility within and between sarcomeres. J. Cell Biol. 193, 785–798.10.1083/jcb.201010099Search in Google Scholar PubMed PubMed Central

Depre, C., Wang, Q., Yan, L., Hedhli, N., Peter, P., Chen, L., Hong, C., Hittinger, L., Ghaleh, B., Sadoshima, J., et al. (2006). Activation of the cardiac proteasome during pressure overload promotes ventricular hypertrophy. Circulation 114, 1821–1828.10.1161/CIRCULATIONAHA.106.637827Search in Google Scholar PubMed

Diehl, F., Brown, M.A., van Amerongen, M.J., Novoyatleva, T., Wietelmann, A., Harriss, J., Ferrazzi, F., Böttger, T., Harvey, R.P., Tucker, P.W., et al. (2010). Cardiac deletion of Smyd2 is dispensable for mouse heart development. PLoS One 5, e9748.10.1371/journal.pone.0009748Search in Google Scholar PubMed PubMed Central

Donlin, L.T., Andresen, C., Just, S., Rudensky, E., Pappas, C.T., Krüger, M., Jacobs, E.Y., Unger, A., Zieseniss, A., Dobenecker, M.W., et al. (2012). Smyd2 controls cytoplasmic lysine methylation of Hsp90 and myofilament organization. Genes Dev. 26, 114–119.10.1101/gad.177758.111Search in Google Scholar PubMed PubMed Central

Ehler, E. and Gautel, M. (2008). The sarcomere and sarcomerogenesis. Adv. Exp. Med. Biol. 642, 1–14.10.1007/978-0-387-84847-1_1Search in Google Scholar PubMed

Fareed, M.U., Evenson, A.R., Wie, W., Menconi, M., Poylin, V., Petkova, V., Pignol, B., and Hasselgren, P.O. (2006). Treatment of rats with calpain inhibitors prevents sepsis-induced muscle proteolysis independent of atrogin-1/MAFbx and MuRF1 expression. Am. J. Physiol. Regul. Integr. Comp. Physiol. 290, R1589–R1597.10.1152/ajpregu.00668.2005Search in Google Scholar PubMed

Fischer, D., Matten, J., Reimann, J., Bönnemann, C., and Schröder, R. (2002). Expression, localization and functional divergence of alphaB-crystallin and heat shock protein 27 in core myopathies and neurogenic atrophy. Acta Neuropathol. 104, 297–304.10.1007/s00401-002-0559-zSearch in Google Scholar PubMed

Freiburg, A., Trombitas, K., Hell, W., Cazorla, O., Fougerousse, F., Centner, T., Kolmerer, B., Witt, C., Beckmann, J.S., Gregorio, C.C., et al. (2000). Series of exon-skipping events in the elastic spring region of titin as the structural basis for myofibrillar elastic diversity. Circ. Res. 86, 1114–1121.10.1161/01.RES.86.11.1114Search in Google Scholar PubMed

Frey, N., Richardson, J.A., and Olson, E.N. (2000). Calsarcins, a novel family of sarcomeric calcineurin-binding proteins. Proc. Natl. Acad. Sci. USA 97, 14632–14637.10.1073/pnas.260501097Search in Google Scholar PubMed PubMed Central

Gao, C.Q., Sawicki, G., Suarez-Pinzon, W.L., Csont, T., Wozniak, M., Ferdinandy, P., and Schulz, R. (2003). Matrix metalloproteinase-2 mediates cytokine-induced myocardial contractile dysfunction. Cardiovasc. Res. 57, 426–433.10.1016/S0008-6363(02)00719-8Search in Google Scholar

Gautel, M. (2011). Cytoskeletal protein kinases: titin and its relations in mechanosensing. Pflüger’s Arch. 462, 119–134.10.1007/s00424-011-0946-1Search in Google Scholar PubMed PubMed Central

Golenhofen, N., Htun, P., Ness, W., Koob, R., Schaper, W., and Drenckhahn, D. (1999). Binding of the stress protein alpha B-crystallin to cardiac myofibrils correlates with the degree of myocardial damage during ischemia/reperfusion in vivo. J. Mol. Cell. Cardiol. 31, 569–580.10.1006/jmcc.1998.0892Search in Google Scholar PubMed

Golenhofen, N., Arbeiter, A., Koob, R., and Drenckhahn, D. (2002). Ischemia-induced association of the stress protein alpha B-crystallin with I-band portion of cardiac titin. J. Mol. Cell. Cardiol. 34, 309–319.10.1006/jmcc.2001.1513Search in Google Scholar PubMed

Granzier, H.L. and Labeit, S. (2004). The giant protein titin: a major player in myocardial mechanics, signaling, and disease. Circ. Res. 94, 284–295.10.1161/01.RES.0000117769.88862.F8Search in Google Scholar PubMed

Gräter, F., Shen, J., Jiang, H., Gautel, M., and Grubmuller, H. (2005). Mechanically induced titin kinase activation studied by force-probe molecular dynamics simulations. Biophys. J. 88, 790–804.10.1529/biophysj.104.052423Search in Google Scholar PubMed PubMed Central

Gregorio, C.C., Trombitás, K., Centner, T., Kolmerer, B., Stier, G., Kunke, K., Suzuki, K., Obermayr, F., Herrmann, B., Granzier, H., et al. (1998). The NH2 terminus of titin spans the Z-disc: its interaction with a novel 19-kD ligand (T-cap) is required for sarcomeric integrity. J. Cell Biol. 143, 1013–1027.10.1083/jcb.143.4.1013Search in Google Scholar PubMed PubMed Central

Gregorio, C.C., Perry, C.N., and McElhinny, A.S. (2005). Functional properties of the titin/connectin-associated proteins, the muscle-specific RING-finger proteins (MURFs), in striated muscle. J. Muscle Res. Cell Motil. 26, 389–400.10.1007/s10974-005-9021-xSearch in Google Scholar PubMed

Guo, W., Schafer, S., Greaser, M.L., Radke, M.H., Liss, M., Govindarajan, T., Maatz, H., Schulz, H., Li, S., Parrish, A.M., et al. (2012). RBM20, a gene for hereditary cardiomyopathy, regulates titin splicing. Nat. Med. 18, 766–773.10.1038/nm.2693Search in Google Scholar PubMed PubMed Central

Hamdani, N., Bishu, K.G., von Frieling-Salewsky, M., Redfield, M.M., and Linke, W.A. (2013a). Deranged myofilament phosphorylation and function in experimental heart failure with preserved ejection fraction. Cardiovasc. Res. 97, 464–471.10.1093/cvr/cvs353Search in Google Scholar PubMed

Hamdani, N., Krysiak, J., Kreusser, M.M., Neef, S., Dos Remedios, C.G., Maier, L.S., Krüger, M., Backs, J., and Linke, W.A. (2013b). Crucial role for Ca2(+)/calmodulin-dependent protein kinase-II in regulating diastolic stress of normal and failing hearts via titin phosphorylation. Circ. Res. 112, 664–674.10.1161/CIRCRESAHA.111.300105Search in Google Scholar PubMed

Hasselgren, P.O. and Fischer, J.E. (2001). Muscle cachexia: current concepts of intracellular mechanisms and molecular regulation. Ann. Surg. 233, 9–17.10.1097/00000658-200101000-00003Search in Google Scholar PubMed PubMed Central

Hayashi, T., Arimura, T., Itoh-Satoh, M., Ueda, K., Hohda, S., Inagaki, N., Takahashi, M., Hori, H., Yasunami, M., Nishi, H., et al. (2004). Tcap gene mutations in hypertrophic cardiomyopathy and dilated cardiomyopathy. J. Am. Coll. Cardiol. 44, 2192–2201.10.1016/j.jacc.2004.08.058Search in Google Scholar PubMed

Hayashi, C., Ono, Y., Doi, N., Kitamura, F., Tagami, M., Mineki, R., Arai, T., Taguchi, H., Yanagida, M., Hirner, S., et al. (2008). Multiple molecular interactions implicate the connectin/titin N2A region as a modulating scaffold for p94/calpain 3 activity in skeletal muscle. J. Biol. Chem. 283, 14801–14814.10.1074/jbc.M708262200Search in Google Scholar PubMed

Hein, S., Arnon, E., Kostin, S., Schönburg, M., Elsässer, A., Polyakova, V., Bauer, E.P., Klövekorn, W.P., and Schaper, J. (2003). Progression from compensated hypertrophy to failure in the pressure-overloaded human heart: structural deterioration and compensatory mechanisms. Circulation 107, 984–991.10.1161/01.CIR.0000051865.66123.B7Search in Google Scholar PubMed

Heineke, J., Ruetten, H., Willenbockel, C., Gross, S.C., Naguib, M., Schaefer, A., Kempf, T., Hilfiker-Kleiner, D., Caroni, P., Kraft, T., et al. (2005). Attenuation of cardiac remodeling after myocardial infarction by muscle LIM protein-calcineurin signaling at the sarcomeric Z-disc. Proc. Natl. Acad. Sci. USA 102, 1655–1660.10.1073/pnas.0405488102Search in Google Scholar PubMed PubMed Central

Herrmann, J., Ciechanover, A., Lerman, L.O., and Lerman, A. (2004). The ubiquitin–proteasome system in cardiovascular diseases – a hypothesis extended. Cardiovasc. Res. 61, 11–21.10.1016/j.cardiores.2003.09.033Search in Google Scholar PubMed

Hidalgo, C., Hudson, B., Bogomolovas, J., Zhu, Y., Anderson, B., Greaser, M., Labeit, S., and Granzier, H. (2009). PKC phosphorylation of titin’s PEVK element: a novel and conserved pathway for modulating myocardial stiffness. Circ. Res. 105, 631–638.10.1161/CIRCRESAHA.109.198465Search in Google Scholar PubMed PubMed Central

Hidalgo, C.G., Chung, C.S., Saripalli, C., Methawasin, M., Hutchinson, K.R., Tsaprailis, G., Labeit, S., Mattiazzi, A., and Granzier, H.L. (2013). The multifunctional Ca(2+)/calmodulin-dependent protein kinase II delta (CaMKIIδ) phosphorylates cardiac titin’s spring elements. J. Mol. Cell Cardiol. 54, 90–97.10.1016/j.yjmcc.2012.11.012Search in Google Scholar PubMed PubMed Central

Hojayev, B., Rothermel, B.A., Gillette, T.G., and Hill, J.A. (2012). FHL2 binds calcineurin and represses pathological cardiac growth. Mol. Cell Biol. 32, 4025–4034.10.1128/MCB.05948-11Search in Google Scholar PubMed PubMed Central

Horwitz, J. (2003). Alpha-crystallin. Exp. Eye. Res. 76, 145–153.10.1016/S0014-4835(02)00278-6Search in Google Scholar

Hunter, J.J. and Chien, K.R. (1999). Signaling pathways for cardiac hypertrophy and failure. N. Engl. J. Med. 341, 1276–1283.10.1056/NEJM199910213411706Search in Google Scholar PubMed

Kato, K., Shinohara, H., Kurobe, N., Inaguma, Y., Shimizu, K., and Ohshima, K. (1991). Tissue distribution and developmental profiles of immunoreactive alpha B crystallin in the rat determined with a sensitive immunoassay system. Biochim. Biophys. Acta. 1074, 201–208.10.1016/0304-4165(91)90062-LSearch in Google Scholar PubMed

Knöll, R., Hoshijima, M., Hoffman, H.M., Person, V., Lorenzen-Schmidt, I., Bang, M.L., Hayashi, T., Shiga, N., Yasukawa, H., Schaper, W., et al. (2002). The cardiac mechanical stretch sensor machinery involves a Z disc complex that is defective in a subset of human dilated cardiomyopathy. Cell 111, 943–955.10.1016/S0092-8674(02)01226-6Search in Google Scholar PubMed

Knöll, R., Kostin, S., Klede, S., Savvatis, K., Klinge, L., Stehle, I., Gunkel, S., Kötter, S., Babicz, K., Sohns, M., et al. (2010). A common MLP (muscle LIM protein) variant is associated with cardiomyopathy. Circ. Res. 106, 695–704.10.1161/CIRCRESAHA.109.206243Search in Google Scholar PubMed

Knöll, R., Linke, W.A., Zou, P., Miocic, S., Kostin, S., Buyandelger, B., Ku, C.H., Neef, S., Bug, M., Schäfer, K., et al. (2011). Telethonin deficiency is associated with maladaptation to biomechanical stress in the mammalian heart. Circ. Res. 109, 758–769.10.1161/CIRCRESAHA.111.245787Search in Google Scholar PubMed PubMed Central

Kötter, S., Gout, L., von Frieling-Salewsky, M., Müller, A.E., Helling, S., Marcus, K., dos Remedios, C., Linke, W.A., and Krüger, M. (2013). Differential changes in titin domain phosphorylation increase myofilament stiffness in failing human hearts. Cardiovasc. Res. 99, 684–656.10.1093/cvr/cvt144Search in Google Scholar PubMed

Kötter, S., Unger, A., Hamdani, N., Lang, P., Vorgerd, M., Nagel-Steeger, L., and Linke, W.A. (2014). Human myozytes are protected from titin aggregation induced stiffening by small heat shock proteins. J. Cell Biol. 204, 187–202.10.1083/jcb.201306077Search in Google Scholar PubMed PubMed Central

Krüger, M. and Linke, W.A. (2011). The giant protein titin: a regulatory node that integrates myocyte signalling pathways. J. Biol. Chem. 286, 9905–9912.10.1074/jbc.R110.173260Search in Google Scholar PubMed PubMed Central

Krüger, M., Kohl, T., and Linke, W. A. (2006). Developmental changes in passive stiffness myofilament Ca2+ sensitivity due to titin and troponin-I isoform switching are not critically triggered by birth. Am. J. Physiol. Heart Circ. Physiol. 291, H496–H506.10.1152/ajpheart.00114.2006Search in Google Scholar PubMed

Krüger, M., Kötter, S., Grützner, A., Lang, P., Andresen, C., Redfield, M.M., Butt, E., dos Remedios, C.G., and Linke, W.A. (2009). Protein kinase G modulates human myocardial passive stiffness by phosphorylation of the titin springs. Circ. Res. 104, 87–94.10.1161/CIRCRESAHA.108.184408Search in Google Scholar PubMed

Labeit, S., Lahmers, S., Burkart, C., Fong, C., McNabb, M., Witt, S., Witt, C., Labeit, D., and Granzier, H. (2006). Expression of distinct classes of titin isoforms in striated and smooth muscles by alternative splicing, and their conserved interaction with filamins. J. Mol. Biol. 362, 664–681.10.1016/j.jmb.2006.07.077Search in Google Scholar PubMed

Lange, S., Auerbach, D., McLoughlin, P., Perriard, E., Schäfer, B.W., Perriard, J.C., and Ehler, E. (2002). Subcellular targeting of metabolic enzymes to titin in heart muscle may be mediated by DRAL/FHL2. J. Cell Sci. 115, 4925–4936.10.1242/jcs.00181Search in Google Scholar PubMed

Lange, S., Xiang, F., Yakovenko, A., Vihola, A., Hackman, P., Rostkova, E., Kristensen, J., Brandmeier, B., Franzen, G., Hedberg, B., et al. (2005). The kinase domain of titin controls muscle gene expression and protein turnover. Science 308, 1599–1603.10.1126/science.1110463Search in Google Scholar PubMed

Lange, S., Ehler, E., and Gautel, M. (2006). From A to Z and back? Multicompartment proteins in the sarcomere. Trends Cell Biol. 16, 11–18.10.1016/j.tcb.2005.11.007Search in Google Scholar PubMed

Lanneau, D., Wettstein, G., Bonniaud, P., and Garrido, C. (2010). Heat shock proteins: cell protection through protein triage. ScientificWorldJournal 10, 1543–1552.10.1100/tsw.2010.152Search in Google Scholar PubMed PubMed Central

Li, H. and Fernandez, J.M. (2003). Mechanical design of the first proximal Ig domain of human cardiac titin revealed by single molecule force spectroscopy. J. Mol. Biol. 334, 75–86.10.1016/j.jmb.2003.09.036Search in Google Scholar PubMed

Li, H.B., Linke, W.A., Oberhauser, A.F., Carrion-Vazquez, M., Kerkviliet, J.G., Lu, H., Marszalek, P.E., and Fernandez, J.M. (2002). Reverse engineering of the giant muscle protein titin. Nature 418, 998–1002.10.1038/nature00938Search in Google Scholar PubMed

Li, J., Horak, K.M., Su, H., Sanbe, A., Robbins, J., and Wang, X. (2011). Enhancement of proteasomal function protects against cardiac proteinopathy and ischemia/reperfusion injury in mice. J. Clin. Invest. 121, 3689–3700.10.1172/JCI45709Search in Google Scholar PubMed PubMed Central

Li, S., Guo, W., Dewey, C.N., and Greaser, M.L. (2013). Rbm20 regulates titin alternative splicing as a splicing repressor. Nucleic Acids Res. 41, 2659–2672.10.1093/nar/gks1362Search in Google Scholar PubMed PubMed Central

Lim, D.S., Robers, R., and Marian, A.J. (2001). Expression profiling of cardiac genes in human hypertrophic cardiomyopathy. Insight into pathogenesis of phenotypes. J. Am. Coll. Cardiol. 38, 1175–1180.10.1016/S0735-1097(01)01509-1Search in Google Scholar

Linke, W.A. and Hamdani, N. (2014). Gigantic business: titin properties and function through thick and thin. Circ. Res. 114, 1052–1068.10.1161/CIRCRESAHA.114.301286Search in Google Scholar PubMed

Linke, W.A. and Krüger, M. (2010). The giant protein titin as an integrator of myocyte signaling pathways. Physiology 25, 186–198.10.1152/physiol.00005.2010Search in Google Scholar PubMed

Linke, W.A., Ivemeyer, M., Olivieri, N., Kolmerer, B., Rüegg, J.C., and Labeit, S. (1996). Towards a molecular understanding of the elasticity of titin. J. Mol. Biol. 261, 62–71.10.1006/jmbi.1996.0441Search in Google Scholar PubMed

Linke, W.A., Rudy, D.E., Centner, T., Gautel, M., Witt, C., Labeit, S., and Gregorio, C.C. (1999). I-band titin in cardiac muscle is a three-element molecular spring and is critical for maintaining thin filament structure. J. Cell Biol. 146, 631–644.10.1083/jcb.146.3.631Search in Google Scholar PubMed PubMed Central

Liu, J., Tang, M., Mestril, R., and Wang, X. (2006). Aberrant protein aggregation is essential for a mutant desmin to impair the proteolytic function of the ubiquitin–proteasome system in cardiomyocytes. J. Mol. Cell. Cardiol. 40, 451–454.10.1016/j.yjmcc.2005.12.011Search in Google Scholar PubMed

Lutsch, G., Vetter, R., Offhauss, U., Wieske, M., Gröne, H.J., Klemenz, R., Schimke, I., Stahl, J., and Benndorf, R. (1997). Abundance and location of the small heat shock proteins HSP25 and alphaB-crystallin in rat and human heart. Circulation 96, 3466–3476.10.1161/01.CIR.96.10.3466Search in Google Scholar PubMed

Makarenko, I., Opitz, C.A., Leake, M.C., Neagoe, C., Kulke, M., Gwathmey, J.K., del Monte, F., Hajjar, R.J., and Linke, W.A. (2004). Passive stiffness changes caused by upregulation of compliant titin isoforms in human dilated cardiomyopathy hearts. Circ. Res. 95, 708–716.10.1161/01.RES.0000143901.37063.2fSearch in Google Scholar PubMed

Mayans, O., van der Ven, P.F., Wilm, M., Mues, A., Young, P., Fürst, D.O., Wilmanns, M., and Gautel, M. (1998). Structural basis for activation of the titin kinase domain during myofibrillogenesis. Nature 395, 863–869.10.1038/27603Search in Google Scholar PubMed

McElhinny, A.S., Kakinuma, K., Sorimachi, H., Labeit, S., and Gregorio, C.C. (2002). Muscle-specific RING finger-1 interacts with titin to regulate sarcomeric M-line and thick filament structure and may have nuclear functions via its interaction with glucocorticoid modulatory element binding protein-1. J. Cell Biol. 157, 125–136.10.1083/jcb.200108089Search in Google Scholar PubMed PubMed Central

Mchaourab, H.S., Berengian, A.R., and Koteiche, H.A. (1997). Site-directed spin-labeling study of the structure and subunit interactions along a conserved sequence in the alpha-crystallin domain of heatshock protein 27. Evidence of a conserved subunit interface. Biochemistry 36, 14627–14634.10.1021/bi971700sSearch in Google Scholar PubMed

McKenna, N., Meigs, J.B., and Wang, Y.L. (1985). Identical distribution of fluorescently labeled brain and muscle actins in living cardiac fibroblasts and myocytes. J. Cell Biol. 100, 292–296.10.1083/jcb.100.1.292Search in Google Scholar PubMed PubMed Central

Mebratu, Y. and Tesfaigzi, Y. (2009). How ERK1/2 activation controls cell proliferationand cell death: is subcellular localization the answer? Cell Cycle 8, 1168–1175.10.4161/cc.8.8.8147Search in Google Scholar PubMed PubMed Central

Meiners, S., Dreger, H., Fechner, M., Bieler, S., Rother, W., Günther, C., Baumann, G., Stangl, V., and Stangl, K. (2008). Suppression of cardiomyocyte hypertrophy by inhibition of the ubiquitin–proteasome system. Hypertension 51, 302–308.10.1161/HYPERTENSIONAHA.107.097816Search in Google Scholar PubMed

Miller, M.K., Granzier, H., Ehler, E., and Gregorio, C.C. (2004). The sensitive giant: the role of titin-based stretch sensing complexes in the heart. Trends Cell Biol. 14, 119–126.10.1016/j.tcb.2004.01.003Search in Google Scholar PubMed

Mittal, B., Sanger, J.M., and Sanger, J.W. (1987). Visualization of myosin in living cells. J. Cell Biol. 105, 1753–1760.10.1083/jcb.105.4.1753Search in Google Scholar PubMed PubMed Central

Moreira, E.S., Wiltshire, T.J., Faulkner, G., Nilforoushan, A., Vainzof, M., Suzuki, O.T., Valle, G., Reeves, R., Zatz, M., Passos-Bueno, M.R., et al. (2000). Limb-girdle muscular dystrophy type 2G is caused by mutations in the gene encoding the sarcomeric protein telethonin. Nat. Genet. 24, 163–166.10.1038/72822Search in Google Scholar PubMed

Nagueh, S.F., Shah, G., Wu, Y., Torre-Amione, G., King, N.M., Lahmers, S., Witt, C.C., Becker, K., Labeit, S., and Granzier, H.L. (2004). Altered titin expression, myocardial stiffness, and left ventricular function in patients with dilated cardiomyopathy. Circulation 110, 155–162.10.1161/01.CIR.0000135591.37759.AFSearch in Google Scholar PubMed

Neagoe, C., Kulke, M., del Monte, F., Gwathmey, J.K., de Tombe, P.P., Hajjar, R.J., and Linke, W.A. (2002). Titin isoform switch in ischemic human heart disease. Circulation 106, 1333–1341.10.1161/01.CIR.0000029803.93022.93Search in Google Scholar PubMed

Neagoe, C., Opitz, C.A., Makarenko, I., and Linke, W.A. (2003). Gigantic variety: expression patterns of titin isoforms in striated muscles and consequences for myofibrillar passive stiffness. J. Muscle Res. Cell Motil. 24, 175–189.10.1023/A:1026053530766Search in Google Scholar PubMed

Olson, E.N. and Williams, R.S. (2000). Calcineurin signaling and muscle remodeling. Cell 101, 689–692.10.1016/S0092-8674(00)80880-6Search in Google Scholar

Opitz, C.A., Leake, M.C., Makarenko, I., Benes, V., and Linke, W.A. (2004). Developmentally regulated switching of titin size alters myofibrillar stiffness in the perinatal heart. Circ. Res. 94, 967–975.10.1161/01.RES.0000124301.48193.E1Search in Google Scholar PubMed

Pankiv, S., Clausen, T.H., Lamark. T., Brech. A., Bruun. J.A., Outzen, H., Overvatn, A., Bjorkoy, G., and Johansen, T. (2007). p62/SQSTM1 binds directly to Atg8/LC3 to facilitate degradation of ubiqzitinated protein aggregates by autophagy. J. Biol. Chem. 282, 24131–24145.10.1074/jbc.M702824200Search in Google Scholar PubMed

Paulsen, G., Lauritzen, F., Bayer, M.L., Kalhovde, J.M., Ugelstad, I., Owe, S.G., Hallén, J., Bergersen, L.H., and Raastad, T. (2009). Subcellular movement and expression of HSP27, αB-crystallin, and HSP70 after two bouts of eccentric exercise in humans. J. Appl. Physiol. 107, 570–582.10.1152/japplphysiol.00209.2009Search in Google Scholar PubMed

Pizon, V., Iakovenko, A., Van Der Ven, P.F., Kelly, R., Fatu, C., Fürst, D.O., Karsenti, E., and Gautel, M. (2002). Transient association of titin and myosin with microtubules in nascent myofibrils directed by the MURF2 RING-finger protein. J. Cell Sci. 115, 4469–4482.10.1242/jcs.00131Search in Google Scholar PubMed

Powell, S.R. and Divald, A. (2010). The ubiquitin-proteasome system in myocardial ischaemia and preconditioning. Cardiovasc. Res. 85, 303–311.10.1093/cvr/cvp321Search in Google Scholar PubMed PubMed Central

Prado, L.G., Makarenko, I., Andresen, C., Krüger, M., Opitz, C.A., and Linke, W.A. (2005). Isoform diversity of giant proteins in relation to passive and active contractile properties of rabbit skeletal muscles. J. Gen. Physiol. 126, 461–480.10.1085/jgp.200509364Search in Google Scholar PubMed PubMed Central

Predmore, J.M., Wang, P., Davis, F., Bartolone, S., Westfall, M.V., Dyke, D.B., Pagani, F., Powell, S.R., and Day, S.M. (2010). Ubiquitin proteasome dysfunction in human hypertrophic and dilated cardiomyopathies. Circulation 121, 997–1004.10.1161/CIRCULATIONAHA.109.904557Search in Google Scholar PubMed PubMed Central

Raskin, A., Lange, S., Banares, K., Lyon, R.C., Zieseniss, A., Lee, L.K., Yamazaki, K.G., Granzier, H.L., Gregorio, C.C., McCulloch, A.D., et al. (2012). A novel mechanism involving four-and-a half LIM domain protein-1 and extracellular signal-regulated kinase-2 regulates titin phosphorylation and mechanics. J. Biol. Chem. 287, 29273–29284.10.1074/jbc.M112.372839Search in Google Scholar PubMed PubMed Central

Raynaud, F., Fernandez, E., Coulis, G., Aubry, L., Vignon, X., Bleimling, N., Gautel, M., Benyamin, Y., and Ouali, A. (2005). Calpain 1-titin interactions concentrate calpain 1 in the Z-band edges and in the N2-line region within the skeletal myofibril. FEBS Lett. 272, 2578–2590.10.1111/j.1742-4658.2005.04683.xSearch in Google Scholar PubMed

Razeghi, P., Baskin, K.K., Sharma, S., Young, M.E., Stepkowski, S., Essop, M.F., and Taegtmeyer, H. (2006). Atrophy, hypertrophy, and hypoxemia induce transcriptional regulators of the ubiquitin proteasome system in the rat heart. Biochem. Biophys. Res. Commun. 342, 361–364.10.1016/j.bbrc.2006.01.163Search in Google Scholar PubMed

Rief, M., Gautel, M., Schemmel, A., and Gaub, H.E. (1998). The mechanical stability of immunoglobulin and fibronectin III domains in the muscle protein titin measured by atomic force microscopy. Biophys. J. 75, 3008–3014.10.1016/S0006-3495(98)77741-0Search in Google Scholar PubMed PubMed Central

Sanger, J.M., Mittal, B., Pochapin, M.B., and Sanger, J.W. (1986). Myofibrillogenesis in living cells microinjected with fluorescently labeled alpha-actinin. J. Cell Biol. 102, 2053–2066.10.1083/jcb.102.6.2053Search in Google Scholar PubMed PubMed Central

Sawicki, G., Leon, H., Sawicka, J., Sariahmetoglu, M., Schulze, C.J., Scott, P.G., Szczesna-Cordary, D., and Schulz, R. (2005). Degradation of myosin light chain in isolated rat hearts subjected to ischemia-reperfusion injury: a new intracellular target for matrix metalloproteinase-2. Circulation 112, 544–552.10.1161/CIRCULATIONAHA.104.531616Search in Google Scholar PubMed

Schlossarek, S. and Carrier, L. (2011). The ubiquitin-proteasome system in cardiomyopathies. Curr. Opin. Cardiol. 26, 190–195.10.1097/HCO.0b013e32834598feSearch in Google Scholar PubMed

Scholl, F.A., McLoughlin, P., Ehler, E., De Giovanni, C., and Schäfer, B.W. (2000). DRAL is a p53-responsive gene whose four-and-a-half LIM domain protein product induces apoptosis. J. Cell Biol. 151, 495–506.10.1083/jcb.151.3.495Search in Google Scholar PubMed PubMed Central

Seibenhener, M.L., Geetha, T., and Wooten, M.W. (2007). Sequestosome1/p62 – more than just a scaffold. FEBS Lett. 581, 175–179.10.1016/j.febslet.2006.12.027Search in Google Scholar PubMed PubMed Central

Sheikh, F., Raskin, A., Chu, P.H., Lange, S., Domenighetti, A.A., Zheng, M., Liang, X., Zhang, T., Yajima, T., Gu, Y., et al. (2008). An FHL1-containing complex within the cardiomyocyte sarcomere mediates hypertrophic biomechanical stress responses in mice. J. Clin. Invest. 118, 3870–3880.10.1172/JCI34472Search in Google Scholar PubMed PubMed Central

Sung, M.M., Schulz, C.G., Wang, W., Sawicki, G., Bautista-Lopez, N.L., and Schulz, R. (2007). Matrix metalloproteinase-2 degrades the cytoskeletal protein α-actinin in peroxynitrite mediated myocardial injury. J. Mol. Cell Cardiol. 43, 429–436.10.1016/j.yjmcc.2007.07.055Search in Google Scholar PubMed

Tanoue, T., Adachi, M., Moriguchi, T., and Nishida, E. (2000). A conserved docking motif in MAP kinases common to substrates, activators and regulators. Nat. Cell Biol. 2, 110–116.10.1038/35000065Search in Google Scholar PubMed

Trombitas, K., Greaser, M., Labeit, S., Jin, J.P., Kellermayer, M., Helmes, M., and Granzier, H. (1998). Titin extensibility in situ: entropic elasticity of permanently folded and permanently unfolded molecular segments. J. Cell Biol. 140, 853–859.10.1083/jcb.140.4.853Search in Google Scholar PubMed PubMed Central

Tskhovrebova, L. and Trinick, J. (2004). Properties of titin immunoglobulin and fibronectin-3 domains. J. Biol. Chem. 279, 46351–46354.10.1074/jbc.R400023200Search in Google Scholar PubMed

Tskhovrebova, L. and Trinick, J. (2010). Roles of titin in the structure and elasticity of the sarcomere. J. Biomed. Biotechnol. 2010, 612482.10.1155/2010/612482Search in Google Scholar PubMed PubMed Central

Tsukamoto, O., Minamino, T., Okada, K., Shintani, Y., Takashima, S., Kato, H., Liao, Y., Okazaki, H., Asai, M., Hirata, A., et al. (2006). Depression of proteasome activities during the progression of cardiac dysfunction in pressure-overloaded heart of mice. Biochem. Biophys. Res. Commun. 340, 1125–1133.10.1016/j.bbrc.2005.12.120Search in Google Scholar PubMed

van de Klundert, F.A., Gijsen, M.L., van den IJssel, P.R., Snoeckx, L.H., and de Jong, W.W. (1998). αB-crystallin and hsp25 in neonatal cardiac cells—differences in cellular localization under stress conditions. Eur. J. Cell Biol. 75, 38–45.10.1016/S0171-9335(98)80044-7Search in Google Scholar PubMed

Voelkel, T., Andresen, C., Unger, A., Just, S., Rottbauer, W., and Linke, W.A. (2013). Lysine methyltransferase Smyd2 regulates Hsp90-mediated protection of the sarcomeric titin springs and cardiac function. Biochim. Biophys. Acta 1833, 812–822.10.1016/j.bbamcr.2012.09.012Search in Google Scholar PubMed

Vos, M.J., Hageman, J., Carra, S., and Kampinga, H.H. (2008). Structural and functional diversities between members of the human HSPB, HSPH, HSPA, and DNAJ chaperone families. Biochemistry 47, 7001–7011.10.1021/bi800639zSearch in Google Scholar PubMed

Wang, X. and Robbins, J. (2006). Heart failure and protein quality control. Circ Res. 99, 1315–1328.10.1161/01.RES.0000252342.61447.a2Search in Google Scholar PubMed

Wang, W., Schulze, C.J., Suarez-Pinzon, W.L., Dyck, J.R., Sawicki, G., and Schulz, R. (2002). Intracellular action of matrix metalloproteinase-2 accounts for acute myocardial ischemia and reperfusion injury. Circulation 106, 1543–1549.10.1161/01.CIR.0000028818.33488.7BSearch in Google Scholar PubMed

Watanabe, K., Muhle-Goll, C., Kellermayer, M.S.Z., Labeit, S., and Granzier, H. (2002). Different molecular mechanics displayed by titin’s constitutively and differentially expressed tandem Ig segments. J. Struct. Biol. 137, 248–258.10.1006/jsbi.2002.4458Search in Google Scholar PubMed

Waters, S., Marchbank, K., Solomon, E., Whitehouse, C., and Gautel, M. (2009). Interactions with LC3 and polyubiquitin chains link nbr1 to autophagic protein turnover. FEBS Lett. 583, 1846–1852.10.1016/j.febslet.2009.04.049Search in Google Scholar PubMed

Watkins, S.J., Borthwick, G.M., and Arthur, H.M. (2011). The H9C2 cell line and primary neonatal cardiomyocyte cells show similar hypertrophic responses in vitro. In Vitro Cell Dev. Biol. Anim. 47, 125–131.10.1007/s11626-010-9368-1Search in Google Scholar PubMed

Weinert, S., Bergmann, N., Luo, X., Erdmann, B., and Gotthardt, M. (2006). M line-deficient titin causes cardiac lethality through impaired maturation of the sarcomere. J. Cell Biol. 173, 559–570.10.1083/jcb.200601014Search in Google Scholar PubMed PubMed Central

Williams, A.B., Decourten-Myers, G.M., Fischer, J.E., Luo, G., Sun, X., and Hasselgren, P.O. (1999). Sepsis stimulates release of myofilaments in skeletal muscle by a calcium-dependent mechanism. FASEB J. 13, 1435–1443.10.1096/fasebj.13.11.1435Search in Google Scholar PubMed

Witt, S.H., Granzier. H., Witt, C.C., and Labeit, S. (2005). MURF-1 and MURF-2 target a specific subset of myofibrillar proteins redundantly: towards understanding MURF-dependent muscle ubiquitination. J. Mol. Biol. 350, 713–722.10.1016/j.jmb.2005.05.021Search in Google Scholar PubMed

Witt, C.C., Burkart, C., Labeit, D., McNabb, M., Wu, Y., Granzier, H., and Labeit, S. (2006). Nebulin regulates thin filament length, contractility, and Z-disk structure in vivo. EMBO J. 25, 3843–3855.10.1038/sj.emboj.7601242Search in Google Scholar PubMed PubMed Central

Wohlschlaeger, J., Sixt, S.U., Stoeppler, T., Schmitz, K.J., Levkau, B., Tsagakis, K., Vahlhaus, C., Schmid, C., Peters, J., Schmid, K.W., et al. (2010). Ventricular unloading is associated with increased 20s proteasome protein expression in the myocardium. J. Heart Lung Transplant. 29, 125–132.10.1016/j.healun.2009.07.022Search in Google Scholar PubMed

Yamasaki, R., Wu, Y., McNabb, M., Greaser, M., Labeit, S., and Granzier, H. (2002). Protein kinase A phosphorylates titin’s cardiac-specific N2B domain and reduces passive tension in rat cardiac myocytes. Circ. Res. 90, 1181–1188.10.1161/01.RES.0000021115.24712.99Search in Google Scholar PubMed

Zhu, Y., Bogomolovas, J., Labeit, S., and Granzier, H. (2009). Single molecule force spectroscopy of the cardiac titin N2B element: effects of the molecular chaperone αB-crystallin with disease-causing mutations. J. Biol. Chem. 284, 13914–13923.10.1074/jbc.M809743200Search in Google Scholar PubMed PubMed Central

Zou, P., Pinotsis, N., Lange, S., Song, Y.H., Popov, A., Mavridis, L., Mayans, O.M., Gautel, M., and Wilmanns, M. (2006). Palindromic assembly of the giant muscle protein titin in the sarcomeric Z-disk. Nature 439, 229–233.10.1038/nature04343Search in Google Scholar PubMed

Received: 2014-3-30
Accepted: 2014-4-23
Published Online: 2014-9-2
Published in Print: 2014-11-1

©2014 by De Gruyter

Downloaded on 2.6.2024 from https://www.degruyter.com/document/doi/10.1515/hsz-2014-0178/html
Scroll to top button