Hostname: page-component-848d4c4894-wg55d Total loading time: 0 Render date: 2024-05-25T19:18:24.995Z Has data issue: false hasContentIssue false

Myofilament-associated proteins with intrinsic disorder (MAPIDs) and their resolution by computational modeling

Published online by Cambridge University Press:  11 January 2023

Bin Sun
Affiliation:
Research Center for Pharmacoinformatics (The State-Province Key Laboratories of Biomedicine-Pharmaceutics of China), Department of Medicinal Chemistry and Natural Medicine Chemistry, College of Pharmacy, Harbin Medical University, Harbin 150081, China
Peter M. Kekenes-Huskey*
Affiliation:
Department of Cell and Molecular Physiology, Loyola University Chicago, IL 60153, USA
*
Author for correspondence: Peter M. Kekenes-Huskey, E-mail: pkekeneshuskey@luc.edu
Rights & Permissions [Opens in a new window]

Abstract

The cardiac sarcomere is a cellular structure in the heart that enables muscle cells to contract. Dozens of proteins belong to the cardiac sarcomere, which work in tandem to generate force and adapt to demands on cardiac output. Intriguingly, the majority of these proteins have significant intrinsic disorder that contributes to their functions, yet the biophysics of these intrinsically disordered regions (IDRs) have been characterized in limited detail. In this review, we first enumerate these myofilament-associated proteins with intrinsic disorder (MAPIDs) and recent biophysical studies to characterize their IDRs. We secondly summarize the biophysics governing IDR properties and the state-of-the-art in computational tools toward MAPID identification and characterization of their conformation ensembles. We conclude with an overview of future computational approaches toward broadening the understanding of intrinsic disorder in the cardiac sarcomere.

Type
Review
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
Copyright © The Author(s), 2023. Published by Cambridge University Press

Purpose and scope

The contractile cells of the heart rely on myofilament proteins that transduce a chemical trigger, calcium (Ca2+), into mechanical force. The myofilament proteins form macromolecular assemblies that perform diverse structural, functional, and regulatory roles. While the composition of these assemblies and their three-dimensional structures continue to be resolved, a high percentage of myofilament proteins contain intrinsically disordered region (IDR)s that do not easily lend themselves to conventional structure determination techniques, such as X-ray crystallography or nuclear magnetic resonance (NMR) spectroscopy. We introduce the term myofilament-associated protein with intrinsic disorder (MAPID) to refer to those myofilament proteins that contain IDRs. We also use the term intrinsically disordered protein (IDP) to refer to proteins that are predominantly unfolded to distinguish them from folded proteins with regions of disorder. Studies and reviews to date have largely acknowledged the existence of IDRs in these proteins, though some reports have gone further to examine how IDRs tune filament protein-binding affinities (Uversky et al., Reference Uversky, Shah, Gritsyna, Hitchcock-DeGregori and Kostyukova2011) and contribute to cardiomyopathy (Na et al., Reference Na, Kong, Straight, Pinto and Uversky2016). However, the IDRs in the majority of MAPIDs are not characterized in detail and thus their roles in myofilament function are largely unexplored. As such, determining the structure/function relationships of these MAPIDs is a final frontier in understanding myofilament physiology.

Experimental and computational methods to structurally and functionally characterize IDPs and IDRs of arbitrary origin have exploded in growth in the last decade. Reviews of IDP structure and structure determination have thus grown in popularity in recent years (Gibbs and Showalter, Reference Gibbs and Showalter2015; Schramm et al., Reference Schramm, Bignon, Brocca, Grandori, Santambrogio and Longhi2019), but they have not been interpreted in the context of proteins essential to heart cell contraction. The purpose of this review therefore is to highlight recent advances in computational approaches developed for, or could be applied, to determining the structures and functions of MAPIDs. We divide this review into two parts: (1) A summary of IDRs in a broad ensemble of myofilament proteins (section ‘Myofilament-associated protein with intrinsic disorder (MAPID)s’) and (2) computational modeling techniques that have been, or could be, applied to MAPIDs (section ‘Computational methods for predicting conformation ensembles of isolated MAPIDs’ to section ‘Computational methods for predicting the MAPID co-assembly’). We emphasize the current state-of-the-art in the computational modeling of myofilament IDPs that have been published in the last five years, where possible, although some older studies are included for context. These innovations are introduced in parallel with high-level discussions of complementary experimental techniques.

Part 1: myofilament-associated protein with intrinsic disorder (MAPID)s

In this part, we introduce myofilament proteins and their roles in cardiac function. We next overview the proteins most commonly associated with the myofilament, and the propensity of IDRs in those proteins. Thereafter, we discuss how these IDRs influence myofilament function and dysfunction, as well as prominent challenges in characterizing their properties.

The physiology of myofilament proteins

Molecular function of myofilament proteins

Cardiac contraction is driven by the concerted activity of myofilament proteins that contract the sarcomeres of the cell (see Fig. 1). Although the major protein components of the sarcomere have been identified, the composition of the sarcomere is dynamic (Willis et al., Reference Willis, Schisler, Portbury and Patterson2008). For this reason, myofilament protein isoform expression can vary during development and in response to pathological stimuli (Marston and Redwood, Reference Marston and Redwood2003). Therefore, we limit the scope of this review to the myofilament-associated genes of the adult rat cardiac myofilament reported in Kooij et al. (Reference Kooij, Venkatraman, Kirk, Ubaida-Mohien, Graham, Faber and Van Eyk2014) and depicted in Fig. 1. Myofilament proteins can be loosely divided into those belonging to the thin filament (section ‘MAPIDs of the thin filament’), the thick filament (section ‘MAPIDs of the thick filament’), and Z-disk (section ‘MAPIDs of the Z-disk’). The thick filaments are formed from the intertwining tails of myosin protein dimers (see representative structure in Fig. 2). These myosins bind to actin monomers of the thin filament, upon which energy released by the hydrolysis of adenosine triphosphate (ATP) is used to generate mechanical force. The thin filament comprises approximately 15 actin (ACTC1) monomers, two troponin macromolecules, and two tropomyosin chains that together form the repeating contractile unit of the sarcomere (Yamada et al., Reference Yamada, Namba and Fujii2020). Thin filament proteins primarily sense elevated intracellular Ca2+ following an action potential to unveil binding sites on actin for myosin. The thin filaments of adjacent sarcomeres are joined by proteins that form the Z-disk. In addition to forming a scaffold for myofilaments, Z-disk proteins are subject to, and perform, sundry regulatory roles that help adapt sarcomere force generation to demand. Several proteins including myosin-binding protein C (MyBPC3) and nebulin bridge filaments or link the Z-disk to filaments, which are also discussed in sections ‘MAPIDs of the thin filament’ and ‘MAPIDs of the Z-disk’. Excellent reviews on myofilament proteins and their functions include Russell and Solís (Reference Russell and Solís2021) and others (Sols and Solaro, Reference Sols and Solaro2021), albeit with limited discussion of their intrinsic disorder.

Fig. 1. (a) Schematic illustration of the sarcomere (drawn with BioRender). (b) In this review, we focus on the cardiac proteins proposed in Kooij et al. (Reference Kooij, Venkatraman, Kirk, Ubaida-Mohien, Graham, Faber and Van Eyk2014) with some additional noteworthy examples. For proteins with multiple isoforms, only isoforms with spectra counts (SC) >10 were selected. Proteins with IDR(s) are indicated by red *. Double ** indicates that the IDR(s) has been experimentally confirmed for the gene, while a single * indicates that the confirmation was based on a related isoform or via bioinformatic predictions (see Table 1 for details).

Fig. 2. Core proteins of the thin and thick filament, based on the schematic from Harris et al. (Reference Harris, Lyons and Bezold2011). The thin filament structure PDB 6KN8 was constructed from a cryo-EM study (Yamada et al., Reference Yamada, Namba and Fujii2020). PDB 5TBY was generated from homology modeling. The MyBPC3 structure was predicted by AlphaFold and was downloaded from the UniprotKB database. A 2 nm resolution model of tarantula thick filament was built by fitting atomistic component structures to EM density map (PDB 3DTP (Alamo et al., Reference Alamo, Wriggers, Pinto, Bártoli, Salazar, Zhao, Craig and Padrón2008)).

Contraction begins with the resting sarcomere. In that state, the myosin binding sites on the thin filament are mostly blocked by tropomyosin at resting Ca2+ levels during diastole (ca. 100 nM) (Clapham, Reference Clapham2007). Activation of calcium channels on the plasma membrane and sarcoplasmic reticulum (SR) following depolarization of the cell conducts Ca2+ and thereby rapidly increases the intracellular Ca2+. TnC, a Troponin (Tn) protein, binds to one equivalent of free Ca2+, which exposes a hydrophobic domain on its N-terminus. The exposed domain provides a binding site for the TnI C-terminus. TnC/TnI binding leads to shifts in the positions of TnT and tropomyosin (Tm). As Tm slides along the actin filament (Rynkiewicz et al., Reference Rynkiewicz, Schott, Orzechowski, Lehman and Fischer2015), binding sites on actin for myosin are unveiled. Actin-bound myosin results in what is commonly called a cross-bridge. Following thin filament activation, cross-bridge formation and ATP hydrolysis generate force that contracts sarcomere. Thin filament activation and cross-bridge formation are highly cooperative, that is, the activation of one contractile unit facilitates the activation of its neighbors. Restoration of diastolic calcium level via SR Ca2+ pumps and plasma membrane (PM) ion exchangers ultimately returns the sarcomere to its relaxed state. Many genes beyond those named here (see Fig. 1) couple the Z-disks and filaments, as well as tune the sarcomere's responsiveness to Ca2+, stretch, and external forces (namely those arising from the filling of ventricles and atria).

Sarcomere contraction is tightly regulated on a beat-to-beat basis and dynamically adapts to demands on cardiac output. Rapid regulation is afforded through post-translational modification (PTM) of myofilament proteins that include phosphorylation, oxidation, ubiquitination, acetylation, and methylation among others (Liddy et al., Reference Liddy, White and Cordwell2013; Jin et al., Reference Jin, Diffee, Colman, anderson and Ge2019). To date, phosphorylation is likely the best understood of myofilament PTMs. Many sites are suggested for titin (999 sites), myosin (87 sites for MYH6 and 126 sites for MYH7), TnI (20 sites), and MyBPC3 (35 sites) based on our queries of the PhosphoSitePlus database (Hornbeck et al., Reference Hornbeck, Zhang, Murray, Kornhauser, Latham and Skrzypek2015). Phosphorylation most commonly occurs through the kinases Ca2+/calmodulin-dependent protein kinase II (CaMKII) and protein kinase A (PKA), as well as protein kinase C (PKC) (DeSantiago et al., Reference DeSantiago, Maier and Bers2002; LeWinter, Reference LeWinter2005; Hidalgo et al., Reference Hidalgo, Chung, Saripalli, Methawasin, Hutchinson, Tsaprailis, Labeit, Mattiazzi and Granzier2013). PKA generally increases force by potentiating Ca2+ release and enhances relaxation by decreasing the Ca2+-sensitivity of force generation, while PKC typically opposes these changes (LeWinter, Reference LeWinter2005); CaMKII is implicated in accelerating the rate of relaxation while the heart is pacing rapidly (DeSantiago et al., Reference DeSantiago, Maier and Bers2002). However, these adaptations are sensitive to the specific sites that are phosphorylated. Recent applications of high-throughput mass spectrometry are revealing a host of other PTM modalities in myofilament proteins (Jin et al., Reference Jin, Diffee, Colman, anderson and Ge2019), including acetylation, methylation, oxidation by reactive oxidation species, and conjugation with sundry other biomolecules. Unlike myofilament protein phosphorylation, these PTMs and their impact on contractility are less understood. As will be later discussed, PTMs are common in IDRs of myofilament proteins.

Other regulatory changes include cellular and tissue adaptations to demands on cardiac output that generally occur more slowly than those afforded by PTMs. Hypertrophic adaptations, for instance, result in enlargement of the heart due to physiological drivers (development, exercise, and pregnancy) and pathophysiological (congenital defects, disease and infection, lifestyle or adverse environmental) conditions. The heart may also undergo atrophic changes that reduce the heart size and under pathophysiological conditions, the thinning of cardiac tissue, which altogether decrease cardiac output. These adaptations can entail the up- and down-regulation of myofilament proteins including their isoforms (Bers, Reference Bers2001), and changes in the number and assembly of sarcomeres (Martin and Kirk, Reference Martin and Kirk2020). While important, these modalities of regulation generally impact the number and organization of myofilament proteins, not their intrinsic properties, and are thus beyond the scope of our review.

Cardiac disease remains one of the most prolific causes of death. The majority of etiologies correspond to pathological adaptations to diet or sedentary lifestyle (Forman and Bulwer, Reference Forman and Bulwer2006), although hereditary origins to congenital defects are also common. At the myofilament protein level, cardiac disease can be accompanied by dysregulated contractility, such as altered Ca2+ sensitivity (Messer and Marston, Reference Messer and Marston2014), kinetics of force generation (Belus et al., Reference Belus, Piroddi, Scellini, Tesi, Amati, Girolami, Yacoub, Cecchi, Olivotto and Poggesi2008), maximum contractile force (Crocini and Gotthardt, Reference Crocini and Gotthardt2021), and cooperativity (Ramirez-Correa et al., Reference Ramirez-Correa, Frazier, Zhu, Zhang, Rappold, Kooij, Bedja, Snyder, Lugo-Fagundo, Hariharan, Li, Shen, Gao, Cingolani, Takimoto, Foster and Murphy2015). Genetic causes or susceptibilities can include missense mutations in myofilament proteins or translational defects (Xu et al., Reference Xu, Dewey, Nguyen and Gomes2010; Mazelin et al., Reference Mazelin, Panthu, Nicot, Belotti, Tintignac, Teixeira, Zhang, Risson, Baas, Delaune, Derumeaux, Taillandier, Ohlmann, Ovize, Gangloff and Schaeffer2016). As an example, the cardiac myosin isoforms MYH7 and MYH6 have 58 and 3 disease-associated variants, respectively, and another 247 and 10 variants of unknown significance (VUSs) based on the ClinVar database (Landrum et al., Reference Landrum, Lee, Riley, Jang, Rubinstein, Church and Maglott2013).Footnote 1 A variety of these mutations exhibit loss- or gain-of-function at the protein-level (Moore et al., Reference Moore, Leinwand, Warshaw, Robbins, Seidman and Watkins2012), which stem from impacts on myosin's intrinsic properties or its interactions with other myofilament proteins. Similarly, PTMs within its IDRs can also contribute to myosin dysfunction (Mahmud et al., Reference Mahmud, Dhami, Rans, Liu and Hwang2021). While gain- and loss-of-function phenotypes at the protein level are unlikely to explain all aspects of a pathological phenotype (Ono et al., Reference Ono, Burgess, Schroder, Elayi, anderson, January, Sun, Immadisetty, Kekenes-Huskey and Delisle2020), studies of IDRs are instructive for understanding mechanisms of myofilament dysfunction.

Myofilament proteins and structure determination efforts

Determining the structures of myofilament proteins at atomistic resolution is an important preliminary step in uncovering their functional roles. The prevalence of well-folded myofilament-associated proteins has enabled hundreds of structural studies via X-ray crystallography, NMR, cryo-electron microscopy (cryo-EM), and small-angle X-ray scattering (SAXS). Troponin C (TnC) from chicken was one of the first myofilament proteins whose structure was determined in atomistic detail, initially via crystallography in 1988 (Ka et al., Reference Ka, Rao, Pyzalska, Drendel, Greaser and Sundaralingam1988) and later by NMR (Dvoretsky et al., Reference Dvoretsky, Abusamhadneh, Howarth and Rosevear2002). Structures of the troponin complex, including the complete TnC molecule with fragments of TnI and TnT were more recently resolved via X-ray (PDB: 1J1E) in 2003 (Takeda et al., Reference Takeda, Yamashita, Maeda and Maeda2003). Macromolecular structures of intact filaments or the Z-disk are less common, but have relied on techniques including cryo-EM spectroscopy, SAXS, and computational protein/protein docking techniques (Alamo et al., Reference Alamo, Qi, Wriggers, Pinto, Zhu, Bilbao, Gillilan, Hu and Padron2016; Yamada et al., Reference Yamada, Namba and Fujii2020; Wang et al., Reference Wang, Grange, Wagner, Kho, Gautel and Raunser2021). As an example, a reconstruction of the thin filament was obtained by docking proteins like troponin and tropomyosin to actin filaments, using data collected from cryo-EM (Yamada et al., Reference Yamada, Namba and Fujii2020). Similar approaches were also used for the thick filament (Alamo et al., Reference Alamo, Wriggers, Pinto, Bártoli, Salazar, Zhao, Craig and Padrón2008) and the Z-disk (Wang et al., Reference Wang, Grange, Wagner, Kho, Gautel and Raunser2021). Despite the strengths of these methods in determining the Angstrom-resolution structures of many well-folded myofilament proteins, at most limited details of IDRs are revealed through these approaches. The paucity of IDR information in these structural models therefore leaves a large gap for linking structure to function.

Computational approaches have grown in tandem with experimental techniques to utilize and inform structure determination studies of myofilament proteins. Computer simulations of protein structure, properties, and functions have long served and will continue to play vital roles in elucidating myofilament mechanics and regulation. Early studies relied on descriptions of proteins as static bodies that could interact through steric and long-range interactions (Millman and Irving, Reference Millman and Irving1988). For instance, myofilaments have been described as charged rods with negative electrostatic potentials, from which electrostatic fields within the myofibril could be predicted (Millman and Irving, Reference Millman and Irving1988). Dynamic models of myofilament proteins have largely consisted of molecular dynamics (MD) simulations, coarse-grained simulations, and implicit representations (Kekenes-Huskey et al., Reference Kekenes-Huskey, Liao, Gillette, Hake, Zhang, Michailova, McCulloch and McCammon2013; Lindert et al., Reference Lindert, Cheng, Kekenes-Huskey, Regnier and McCammon2015; Aboelkassem et al., Reference Aboelkassem, McCabe, Huber, Regnier, McCammon and McCulloch2019), which have been made possible through the availability of hundreds of experimentally determined protein structures. These simulations have provided critical insights into the molecular mechanisms underlying sarcomere contraction, its modulation by PTMs, and impacts of missense variants on contractile function. Bosswman and Lindert (Reference Bowman and Lindert2019) offer an excellent review of such applications applied to folded myofilament proteins.

Myofilament-associated protein with intrinsic disorder (MAPID)s

Despite advances made with well-folded, globular proteins, a substantial fraction of the myofilament lacks well-folded structure. Proteins lacking folded structures are referred to as proteins with IDRs when partially folded, or as an IDP if the protein is mostly or completely unfolded. We describe such myofilament proteins as MAPIDs to distinguish them from their well-folded myofilament counterparts. MAPIDs play pivotal roles in myofilament function. One such example is the troponin complex consisting of TnI, TnC, and TnT proteins that triggers muscle contraction after binding Ca2+ (Metskas and Rhoades, Reference Metskas and Rhoades2016). Both TnI and TnT have IDRs that are involved in initiating contraction (Hoffman and Sykes, Reference Hoffman and Sykes2008; Hwang et al., Reference Hwang, Cai, Pineda-Sanabria, Corson and Sykes2014; Johnston et al., Reference Johnston, Landim-Vieira, Marques, Oliveira, Gonzalez-Martinez, Moraes, He, Iqbal, Wilnai, Birk, Zucker, Silva, Chase and Pinto2019). Another example is the behemoth titin protein, which features numerous proline, glutamate, valine and lysine-rich (PEVK, ~28 residue IDR (Linke et al., Reference Linke, Ivemeyer, Mundel, Stockmeier and Kolmerer1998; Ma and Wang, Reference Ma and Wang2003)) repeats that help maintain passive tension in myocytes (Yamasaki et al., Reference Yamasaki, Berri, Wu, Trombitas, McNabb, Kellermayer, Witt, Labeit, Labeit, Greaser and Granzier2001). In fact, IDRs appear to be very common among the proteins of the myofilament. To estimate their propensity, we used the PONDR software ‘PONDR-VLXT’ (Li et al., Reference Li, Romero, Rani, Dunker and Obradovic1999; Romero et al., Reference Romero, Obradovic, Li, Garner, Brown and Dunker2001) to predict IDRs in the sequences of myofilament proteins listed in Fig. 3. PONDR identified that among over 30 myofilament genes that we considered in this work, approximately 42% of the amino acids in the sequences have potential disorder. This number resembles estimates of 30–50% for the entire Eukaryotic proteome (Best, Reference Best2017; Clarke and Pappu, Reference Clarke and Pappu2017).

Fig. 3. (a) The IDP-phase diagram developed by the Pappu lab (Holehouse et al., Reference Holehouse, Das, Ahad, Richardson and Pappu2017), which groups proteins by characteristic disorder including molten, extended, or compact (Uversky, Reference Uversky and Renaud2020) classes have also been proposed based on their charge patterns (R1–R5) (Holehouse et al., Reference Holehouse, Das, Ahad, Richardson and Pappu2017): R1 corresponds to weak polyampholytes and resembles pre-molten globules. R3 signifies strong polyampholytes with a comparable amount of positively and negatively charged residues and is described as hairpins/coils/chimeras. R2 is the boundary between R1 and R3 where coils and pre-molten globules coexist. R4 and R5 are strong polyampholytes like R3, but with dominant negative and positive residues, respectively. IDPs in R4 and R5 are swollen coils. (be) PONDR-VLXT predicted disordered regions in cardiac myofilament proteins. These proteins are categorized into thin/thick filament(s), Z-disk, and miscellaneous. The IDP region is colored red and interlaced with folded regions. The blue line depicts the first and last amino acid and the number is increasing counterclockwise. The numbers in the parentheses present the percentage of predicted IDR residues, ‘pathogenic or likely pathogenic’ mutations, and phosphorylation sites located in the predicted IDP regions, respectively. The structural state estimation of predicted >5 residue IDR regions before (black dots) and after phosphorylation (magenta dots, if PTM site exists in the IDP region) in the IDP-phase diagram were also shown.

In this review, we discuss predominant myofilament proteins identified in the adult rat cardiac myofilament by Kooij et al. (Reference Kooij, Venkatraman, Kirk, Ubaida-Mohien, Graham, Faber and Van Eyk2014) and several additional genes of recent interest including ENH2, MYOT, NEB, MYPN, and LMOD2. We limit our discussions to the major isoforms of these proteins. For those with multiple isoforms, we select the most common based on Kooij et al..Footnote 2 We classify these proteins further using the classification scheme introduced in section ‘Molecular function of myofilament proteins’ and Fig. 1 which defines three regimes: the thin filament, the thick filament, and the Z-disk. For convenience, proteins that are localized to two more regimes, such as MyBPC3 linking the thin and thick filaments, are assigned to a single class. For each gene, we briefly introduce available IDR studies for it, where applicable. Experimental approaches used for the IDR studies are described in part 2. If IDR studies are not reported, we use two sequence-based approaches, the PONDR IDR predictor (Obradovic et al., Reference Obradovic, Peng, Vucetic, Radivojac, Brown and Dunker2003) and IDR state-diagrams (Das and Pappu, Reference Das and Pappu2013; Das et al., Reference Das, Ruff and Pappu2015; Holehouse et al., Reference Holehouse, Das, Ahad, Richardson and Pappu2017) in Fig. 3 to estimate IDR propensity and structure. We also refer to the ClinVar (Landrum et al., Reference Landrum, Lee, Riley, Jang, Rubinstein, Church and Maglott2013) and PhosphoSitePlus (Hornbeck et al., Reference Hornbeck, Zhang, Murray, Kornhauser, Latham and Skrzypek2015) databases for single nucleotide polymorphisms and PTMs within IDRs, respectively.

MAPIDs of the thin filament

We first describe the thin filament proteins, which we divide into two groups: those forming the core of the thin filament and those associated with the thin filament. The core proteins of the thin filament include actin, troponin (TnC, troponin I (TnI), troponin T (TnT)), and Tm. The structures of many of these proteins have been determined down to Angstrom-level resolution, though many of these include IDRs that have not been completely resolved. The core structure of the intact thin filament comprising the troponin complex, actin, and tropomyosin has been resolved by cryo-EM (Yamada et al., Reference Yamada, Namba and Fujii2020) (see Fig. 2).

Core thin filament MAPIDs

Actin (ACTC1) Actin is a 42 kD protein that forms the backbone of the thin filament (Despond and Dawson, Reference Despond and Dawson2018; Frank et al., Reference Frank, Yusuf Rangrez, Friedrich, Dittmann, Stallmeyer, Yadav, Bernt, Schulze-Bahr, Borlepawar, Zimmermann, Peischard, Seebohm, Linke, Baba, Krüger, Unger, Usinger, Frey and Schulze-Bahr2019). In the thin filament, actin has a well-folded structure when co-assembled with troponin and tropomyosin (PDB 6KN8 (Yamada et al., Reference Yamada, Namba and Fujii2020)). Actin-binding proteins participate in actin filament formation (Miao et al., Reference Miao, Tipakornsaowapak, Zheng, Mu and Lewellyn2018). However, experimental evidence suggests that actin does not fold spontaneously without ligand binding or chaperones (Neirynck et al., Reference Neirynck, Waterschoot, Vandekerckhove, Ampe and Rommelaere2006; Turoverov et al., Reference Turoverov, Kuznetsova and Uversky2010). This agrees with bioinformatics studies that suggest actin contains a significant degree of IDR content (Turoverov et al., Reference Turoverov, Kuznetsova and Uversky2010; Povarova et al., Reference Povarova, Uversky, Kuznetsova and Turoverov2014). Within these IDR regions, we identified 42 phosphorylation sites and two disease-associated mutations using the PhosphoSitePlus and ClinVar databases, respectively.

Troponin C (TNNC1) The Tn complex comprises troponin C (TnC), troponin I (TnI), and troponin T (TnT). This complex serves as the central hub on the thin filament that transduces Ca2+ binding to priming the thin filament for myosin binding and ultimately contraction (Marston and Zamora, Reference Marston and Zamora2020). Troponin C (18 kD) has been extensively studied in isolation or in the intact Tn macromolecule with TnI and TnT (Hoffman et al., Reference Hoffman, Blumenschein and Sykes2006; Hoffman and Sykes, Reference Hoffman and Sykes2008; Lindert et al., Reference Lindert, Kekenes-Huskey, Huber, Pierce and McCammon2012a, Reference Lindert, Kekenes-Huskey and McCammon2012b; Zamora et al., Reference Zamora, Papadaki, Messer, Marston and Gould2016; Marques et al., Reference Marques, Parvatiyar, Yang, Oliveira and Pinto2019). The Ca2+ sensor TnC is one of the seemingly few myofilament proteins for which a complete, well-folded structure has been resolved. One of the complete structures for TnC was crystallized as a complex of TnC, TnI, and TnT via X-ray crystallography at 3.3 Å resolution (PDB code:1J1E (Takeda et al., Reference Takeda, Yamashita, Maeda and Maeda2003)). Nonetheless, the linker bridging its N- and C-terminal domains is predicted to be intrinsically disordered (Na et al., Reference Na, Kong, Straight, Pinto and Uversky2016), while predictions using PONDR in Fig. 3 suggest even greater propensity for disorder. We discuss this in greater detail in Fig. S1. Coincidentally, it has few PTMs (4 from PhosphoSitePlus database (Hornbeck et al., Reference Hornbeck, Zhang, Murray, Kornhauser, Latham and Skrzypek2015)) and 10 likely pathogenic variants from the ClinVar database (Landrum et al., Reference Landrum, Lee, Riley, Jang, Rubinstein, Church and Maglott2013).

Troponin I (TNNI3) TnI is a 24 kD protein that binds to a hydrophobic patch on TnC that is exposed following Ca2+ binding (Marston and Zamora, Reference Marston and Zamora2020). TnI binding to TnC primes the thin filament for myosin/actin cross-bridge formation (Marston and Zamora, Reference Marston and Zamora2020). TnI is perhaps the best studied of the IDR-containing proteins that form intact Tn. TnI's N-terminal fragment, which consists of residues M1–H34, is intrinsically disordered and is not represented in troponin crystal structures from Takeda et al. (Reference Takeda, Yamashita, Maeda and Maeda2003). The mobility of the disordered region in the TnI's N-terminal domain is integral to its function (Hoffman et al., Reference Hoffman, Blumenschein and Sykes2006). Using solution NMR spectroscopy, Hwang et al. revealed that this region plays an important role in positioning troponin C for its function (Hwang et al., Reference Hwang, Cai, Pineda-Sanabria, Corson and Sykes2014) and its conformational fluctuations impact Ca2+-regulated myosin binding to the thin filament (Hoffman et al., Reference Hoffman, Blumenschein and Sykes2006; Hoffman and Sykes, Reference Hoffman and Sykes2008). This IDR also harbors three PTM sites (S5/S23/24) and possible cardiac disease-related mutations (Hwang et al., Reference Hwang, Cai, Pineda-Sanabria, Corson and Sykes2014; Metskas and Rhoades, Reference Metskas and Rhoades2016; Na et al., Reference Na, Kong, Straight, Pinto and Uversky2016). In addition, Takeda et al. suggested that TnI's residues E66–R79 are likely disordered because this region folded only when interacting with troponin C (Takeda et al., Reference Takeda, Yamashita, Maeda and Maeda2003). Residues G137–R146 that form its inhibitory peptide are also unresolved (Takeda et al., Reference Takeda, Yamashita, Maeda and Maeda2003). This inhibitory peptide binds to TnC's hydrophobic patch, which is a process that has been the subject of many computational studies in recent years (Lindert et al., Reference Lindert, Kekenes-Huskey and McCammon2012b, Reference Lindert, Cheng, Kekenes-Huskey, Regnier and McCammon2015; Bowman and Lindert, Reference Bowman and Lindert2019). Lastly, the TnI C-terminus (residues I125–S210) is also predicted to be an IDR (Hoffman and Sykes, Reference Hoffman and Sykes2008) and contains sites for PTMs. For instance, a mouse model examining the phosphorylation of S199, which resides in the C-terminal IDR, was found to impair diastolic cardiac function by increasing Ca2+ sensitivity (Li et al., Reference Li, Zhu, Paolocci, Zhang, Takahashi, Okumus, Heravi, Keceli, Ramirez-Correa, Kass and Murphy2017). Nearby, an acetylation mimetic at K132 exhibited accelerated relaxation relative to the native amino acid (Lin et al., Reference Lin, Schmidt, Fritz, Jeong, Cammarato, Foster, Biesiadecki, McKinsey and Woulfe2020). TnI also presents 12 disease-associated variants as reported by the ClinVar database and tens of likely pathogenic mutations associated with cardiomyopathy (Lu et al., Reference Lu, Wu and Morimoto2013).

Troponin T (TNNT2) TnT is a 36 kD protein that regulates muscle contraction by binding to tropomyosin following Ca2+-activation of TnC/TnI (Marston and Zamora, Reference Marston and Zamora2020). TnT has two IDR regions. The first is a ~50 residue linker (approximately residues R158–Q203) between two structured motifs (Deranek et al., Reference Deranek, Baldo, Lynn, Schwartz and Tardiff2022). Cross-linking mass spectroscopy (MS) shows that the binding of TnT's intrinsically disordered C-terminus to TnC contributes to force generation in the myofilament (Johnston et al., Reference Johnston, Landim-Vieira, Marques, Oliveira, Gonzalez-Martinez, Moraes, He, Iqbal, Wilnai, Birk, Zucker, Silva, Chase and Pinto2019). Using Foerster resonance energy transfer (FRET) and molecular dynamics (MD) simulations, the linker's conformational ensembles on the full cardiac thin filament have been elucidated (Deranek et al., Reference Deranek, Baldo, Lynn, Schwartz and Tardiff2022). The second region is the C-terminus of TnT, which has been predicted to be an IDR (Na et al., Reference Na, Kong, Straight, Pinto and Uversky2016), and confirmed by our PONDR results in Fig. 3. PhosphoSitePlus indicates that the C-terminus harbors several PTM sites (T213, S249, Y251, and T294). Phosphorylation of several of these sites is reported to alter cardiac contractility by either reducing Ca2+ sensitivity or ATPase activity (Streng et al., Reference Streng, de Boer, van der Velden, van Dieijen-Visser and Wodzig2013). Similar to TnI, TnT hosts tens of mutations that are linked to cardiomyopathy (Lu et al., Reference Lu, Wu and Morimoto2013).

Tropomyosin (TPM1 and TPM3) The 33 kD tropomyosin isoforms engage TnT to activate the thin filament. Modeling studies to date have targeted the well-folded helices that shift (Rynkiewicz et al., Reference Rynkiewicz, Schott, Orzechowski, Lehman and Fischer2015) along the actin filament (Lehman, Reference Lehman2016) to unveil myosin-binding sites. Tm forms a flexible coiled-coil structure that binds to the thin filament (Singh and Hitchcock-DeGregori, Reference Singh and Hitchcock-DeGregori2003; Yamada et al., Reference Yamada, Namba and Fujii2020). This flexibility is a key modulator of TM's function, as mutations of a highly conserved residue D137L, and a dilated cardiomyopathy (DCM) mutation D230N in α-tropomyosin (TPM), causes a structural rearrangement of its coiled-coil structure that consequently alters its flexibility (Yar et al., Reference Yar, Chowdhury, Davis, Kobayashi, Monasky, Rajan, Wolska, Gaponenko, Kobayashi, Wieczorek and Solaro2013; Lynn et al., Reference Lynn, Tal, Holeman, Jimenez, Strom and Tardiff2017). These changes ultimately impair tropomyosin (TM) function (Yar et al., Reference Yar, Chowdhury, Davis, Kobayashi, Monasky, Rajan, Wolska, Gaponenko, Kobayashi, Wieczorek and Solaro2013; Lynn et al., Reference Lynn, Tal, Holeman, Jimenez, Strom and Tardiff2017). Its N-terminal domain is confirmed to be an IDR by NMR and circular dichroism (CD) studies, which show that this domain gains helical content upon binding to tropomodulin (Kostyukova et al., Reference Kostyukova, Hitchcock-DeGregori and Greenfield2007). This IDR character is believed to explain the inability to resolve the region via cryo-EM in an earlier study (Milligan et al., Reference Milligan, Whittaker and Safer1990). Two PTM sites (S6/S16 for TPM1 and T5/T6 for TPM3) reside within the N-terminal extension as inferred from PhosphoSitePlus (Hornbeck et al., Reference Hornbeck, Zhang, Murray, Kornhauser, Latham and Skrzypek2015). The ClinVar database also indicates several possible pathogenic variants in the N-terminal IDR (K30Δ for TPM1; E3Q, D15N, and M9R for TPM3). In addition, a K15N mutation in TPM1 is reported to change actin's slow-growing (pointed) end dynamics (Colpan et al., Reference Colpan, Ly, Grover, Tolkatchev and Kostyukova2017), which may impact sarcomere assembly.

Thin filament associated MAPIDs

Leiomodin (LMOD2) LMOD2 is a 62 kD protein that helps lengthen the thin filament by driving actin assembly at the filaments' barbed ends (Pappas et al., Reference Pappas, Mayfield, Henderson, Jamilpour, Cover, Hernandez, Hutchinson, Chu, Nam, Valdez, Wong, Granzier and Gregorio2015). While three leiomodin isoforms are known, LMOD2 has the highest expression level in cardiac tissue (Tolkatchev et al., Reference Tolkatchev, Kuruba, Smith, Swain, Smith, Moroz, Williams and Kostyukova2021). A linker connecting three actin binding sites in LMOD2 are likely to be intrinsically disordered, since (1) it is enriched in negatively charged residues and (2) was not resolved in its crystal structure (Tolkatchev et al., Reference Tolkatchev, Kuruba, Smith, Swain, Smith, Moroz, Williams and Kostyukova2021). Immunofluorescence imaging and co-sedimentation experiments showed that this potential IDR facilitates the binding of LMOD2 to the thin filament by displacing bound tropomodulin (Tsukada et al., Reference Tsukada, Pappas, Moroz, Antin, Kostyukova and Gregorio2010; Colpan et al., Reference Colpan, Ly, Grover, Tolkatchev and Kostyukova2017; Tolkatchev et al., Reference Tolkatchev, Kuruba, Smith, Swain, Smith, Moroz, Williams and Kostyukova2021). Three possible pathogenic variants within this IDR region are reported in ClinVar (R513Ter, L415fs, and W398Ter) in addition to PTMs at sites Y369, T384, Y390, T409, S412, S416, T420, and T456 in the PhosphoSitePlus database.

Tropomodulin-1 (TMOD1) Tropomodulin (40 kD) is an actin-binding protein that belongs to the same protein family as leiomodin (Tolkatchev et al., Reference Tolkatchev, Kuruba, Smith, Swain, Smith, Moroz, Williams and Kostyukova2021). TMOD1 regulates actin filament assembly (Boczkowska et al., Reference Boczkowska, Rebowski, Kremneva, Lappalainen and Dominguez2015) and requires tropomyosin (Tm) for its regulatory functions (Kostyukova et al., Reference Kostyukova, Hitchcock-DeGregori and Greenfield2007). The N-terminus of tropomodulin is an IDR based on its susceptibility to proteolysis (Kostyukova et al., Reference Kostyukova, Maeda, Yamauchi, Krieger and Maeda2000) and CD spectroscopy (Kostyukova et al., Reference Kostyukova, Hitchcock-DeGregori and Greenfield2007). The region assumes an alpha-helical configuration, however, when bound to the tropomodulin IDR (Kostyukova et al., Reference Kostyukova, Hitchcock-DeGregori and Greenfield2007). This IDR's dynamic equilibrium exhibits ‘avidity’, in that it favors multiple binding interactions with tropomyosin and actin (Tolkatchev et al., Reference Tolkatchev, Kuruba, Smith, Swain, Smith, Moroz, Williams and Kostyukova2021). The dynamics of the complex binding arrangement was recently examined in a multiscale modeling strategy entailing docking and MD refinement (Tolkatchev et al., Reference Tolkatchev, Kuruba, Smith, Swain, Smith, Moroz, Williams and Kostyukova2021). Putative PTM sites are identified at S2, Y3, Y10, and T23 from PhosphoSitePlus (Hornbeck et al., Reference Hornbeck, Zhang, Murray, Kornhauser, Latham and Skrzypek2015) that may suggest regulatory control of thin filament assembly. Several mutations (A21K/E33V (Moroz et al., Reference Moroz, Guillaud, Desai and Kostyukova2013) and T54E (Dorovkov et al., Reference Dorovkov, Beznosov, Shah, Kotlyanskaya and Kostyukova2008)) in the N-terminal IDR have also been characterized. These mutations are shown to alter TMOD1's binding affinity toward tropomyosin (Moroz et al., Reference Moroz, Guillaud, Desai and Kostyukova2013) and abolish TMOD1's actin capping function (Dorovkov et al., Reference Dorovkov, Beznosov, Shah, Kotlyanskaya and Kostyukova2008). As of yet, no variants of this gene have been reported in ClinVar.

Catenin Alpha 1 (CTNNA1) is a 100 kD mechanosensitive protein that couples the actin cytoskeleton with cadherins of the cell membrane (Vite et al., Reference Vite, Li and Radice2015). CD spectroscopy indicates that its helix bundle E (residues Q260–R360) is intrinsically disordered in the free protein (Hirano et al., Reference Hirano, Amano, Yonemura and Hakoshima2018). This unfolding facilitates binding to vinculin (Hirano et al., Reference Hirano, Amano, Yonemura and Hakoshima2018). Eight PTM sites in the helix E were reported in the PhosphoSitePlus database, in addition to two possible pathogenic variants (E307K and L318S) in ClinVar.

Transglutaminase (TGM2) is a 77 kD protein that catalyzes covalent bonding of glutamine and lysine side chains (Lorand and Graham, Reference Lorand and Graham2003). In the heart, it is implicated in cardiomyocyte development and signaling (Sane et al., Reference Sane, Kontos and Greenberg2007). The enzyme may localize to the thick filament, based on observations of its co-localization with the A-band in cultured, embryonic chicken myoblasts (Kang et al., Reference Kang, Shin, Song, Ha, Chung and Kang1995). TGM2 possesses several IDRs spanning the entire protein as suggested by missing regions within its crystal structures (Pinkas et al., Reference Pinkas, Strop, Brunger and Khosla2007; Kanchan et al., Reference Kanchan, Fuxreiter and Fésüs2015). Bioinformatics studies by Thangaraju et al. indicate that human TGM2 has more IDRs forming short linear motifs (SLIMs) than in other species (Thangaraju et al., Reference Thangaraju, Király, Mótyán, Ambrus, Fuxreiter and Fésüs2017). These SLIMs are important as they enable TGM2 to interact with multiple protein partners, which contributes to its multi-faceted functionality in human (Thangaraju et al., Reference Thangaraju, Király, Mótyán, Ambrus, Fuxreiter and Fésüs2017). PTM sites within potentially disordered loops have been reported in the PhosphoSitePlus database and include Y245, S250, T368, Y369, S415, and S4192. To our knowledge, disease-associated mutations in TGM2 have not yet been reported.

Synaptopodin 2 (SYNPO2) (aka myopodin) is an 118 kD protein involved in actin assembly during myofibril development (Linnemann et al., Reference Linnemann, Vakeel, Bezerra, Orfanos, Djinović-Carugo, Ven, Kirfel and Fürst2013), where it stimulates actin polymerization and aggregation (Chalovich and Schroeter, Reference Chalovich and Schroeter2010). Synaptopodin 2 shares a high sequence identity of about 70% with fesselin. The latter protein has limited secondary structure as measured by CD and its large Stokes radius, which suggest that fesselin, and potentially synaptopodin 2 given its sequence similarity, are unfolded in their native state (Khaymina et al., Reference Khaymina, Kenney, Schroeter and Chalovich2007). Almost 100 PTM sites have been reported for SYNPO2 in PhosphoSitePlus. While we did not identify SYNPO2 mutations in its IDR that are attributed to cardiomyopathy, it has been reported that the reduced expression of SYNPO2 destabilizes myofibrils (Lohanadan et al., Reference Lohanadan, Molt, Dierck, van, Frey, Höhfeld and Fürst2021).

Actin-binding LIM protein 1 (ABLIM1) is an 88 kD protein that traverses the actin cytoskeleton (Roof et al., Reference Roof, Hayes, Adamian, Chishti and Li1997) and links the Z-disk binding domains of titin; the latter engagement is believed to help regulate length-dependent activation of cardiomyocytes (Stachowski-Doll et al., Reference Stachowski-Doll, Papadaki, Martin, Ma, Gong, Shao, Shen, Muntu, Kumar, Perez, Martin, Moravec, Sadayappan, Campbell, Irving and Kirk2022). Although studies of the mammalian ABLIM1 gene's IDRs have not been reported in the literature, the actin-binding plant LIM protein has been experimentally confirmed to have an IDR linker connecting its two LIM domains (Ma and Miao, Reference Ma and Miao2020). Interestingly, this IDR mediates self-aggregation of plant ABLIM proteins and thereby shapes F-actin remodeling in plants (Ma and Miao, Reference Ma and Miao2020). By extension, the homologous region in the cardiac ABLIM1 may also be disordered. We base this suggestion on our PONDR prediction of ABLIM1, which indicates that the residues flanking the first and fourth LIM domains, as well as the C-terminal fragments, are IDRs (Fig. 3). Within these putative IDRs, 92 PTM sites are identified in the PhosphoSitePlus, although no variants have yet been reported in the ClinVar database.

MAPIDs of the thick filament

The thick filament generates force when the thin filament is activated. Myosin is the predominant constituent of the thick filament, while titin and MyBPC3 link the thick filament to the Z-disk (LeWinter and Granzier, Reference LeWinter and Granzier2010) and thin filament (Flashman et al., Reference Flashman, Redwood, MoolmanSmook and Watkins2004), respectively. It is now appreciated that most hypertrophic cardiomyopathy (HCM)-causing mutations are found in thick filament genes MYH7 and MyBPC3 (Xu et al., Reference Xu, Dewey, Nguyen and Gomes2010; Harris et al., Reference Harris, Lyons and Bezold2011), although the thin filament troponin complex presents additional HCM variants (Willott et al., Reference Willott, Gomes, Chang, Parvatiyar, Pinto and Potter2010). DCM mutations have also been identified in MYH7 (Xu et al., Reference Xu, Dewey, Nguyen and Gomes2010), MyBPC3 (Xu et al., Reference Xu, Dewey, Nguyen and Gomes2010), and TTN (Harris et al., Reference Harris, Lyons and Bezold2011).

Myosin (MYH6-7 and MYL2,3,7) The cardiac myosins belong to a super family containing numerous isoforms (Hartman and Spudich, Reference Hartman and Spudich2012). The myosin isoforms comprising the heavy (MYH6 and MYH7) and light (MYL2, MYL3, and MYL7) chain are the work-horses of myofilament contraction, leveraging the hydrolysis of bound ATP to ratchet along exposed actin-binding sites (Spudich et al., Reference Spudich, Finer, Simmons, Ruppel, Patterson and Uyeda1995). The heavy chains are approximately 220 kD while the light chain isoforms are significantly smaller at 20 kD. Our PONDR results in Fig. 3 indicate that there are about 30 putative IDR regions in MYH6/MYH7, and about five in MYL2/MYL3/MYL7. Roughly 130 HCM mutations and 30 DCM mutations are reported in myosin, with most in the MYH7 gene (Carniel et al., Reference Carniel, Taylor, Sinagra, Lenarda, Ku, Fain, Boucek, Cavanaugh, Miocic, Slavov, Graw, Feiger, Zhu, Dao, Ferguson, Bristow and Mestroni2005; Alamo et al., Reference Alamo, Ware, Pinto, Gillilan, Seidman, Seidman and Padron2017; Kim et al., Reference Kim, Park, Lee, Jeon, On, Kim, Choe, Ki, Kim and Kim2020). Consistent with those reports, the ClinVar database reports 346 pathogenic and likely pathogenic variants for these five myosin genes. Of these, 152 reside within its potential IDRs (Fig. 3). The cardiac myosins are also prime targets for PTMs and especially phosphorylation. According to PhosphoSitePlus, nearly 103 of these PTMs fall within potential IDRs for the cardiac myosin genes (Fig. 3). This abundance of PTM sites likely helps regulate thick filament assembly and contraction (Pfitzer, Reference Pfitzer2001). In support of this, it has been reported that blunting myosin light chain 2 phosphorylation leads to abnormal cardiac structure and function in mice (Sanbe et al., Reference Sanbe, Fewell, Gulick, Osinska, Lorenz, Hall, Murray, Kimball, Witt and Robbins1999).

MYH7 is perhaps the most well-studied of the two heavy chains, including several computational studies, as it is the predominant isoform in the human heart (Kelly et al., Reference Kelly, Caleshu, Morales, Buchan, Wolf, Harrison, Cook, Dillon, Garcia, Haverfield, Jongbloed, Macaya, Manrai, Orland, Richard, Spoonamore, Thomas, Thomson, Vincent, Walsh, Watkins, Whiffin, Ingles, Tintelen, Semsarian, Ware, Hershberger and Funke2018). A structural model of the human MYH7 complex, including its myosin light chains, was built from tarantula skeletal muscle thick filaments (Nag et al., Reference Nag, Trivedi, Sarkar, Adhikari, Sunitha, Sutton, Ruppel and Spudich2017). The interactions between protein domains were verified via biochemical assays. Based on this model, some HCM mutations are shown to destabilize the myosin complex, which may explain their detrimental effects on cardiac function (Nag et al., Reference Nag, Trivedi, Sarkar, Adhikari, Sunitha, Sutton, Ruppel and Spudich2017).

One study of note aimed to infer the phenotype for VUSs in well-folded regions of myosin (Toepfer et al., Reference Toepfer, Garfinkel, Venturini, Wakimoto, Repetti, Alamo, Sharma, Agarwal, Ewoldt, Cloonan, Letendre, Lun, Olivotto, Colan, Ashley, Jacoby, Michels, Redwood, Watkins, Day, Staples, Padrón, Chopra, Ho, Chen, Pereira, Seidman and Seidman2020). This study indicated that known pathogenic variants disturb myosin's functional conformation dynamics through altering the myosin head domain's interactions. The proximity of five myosin VUSs to these pathogenic variants was found to correlate with clinical phenotypes; namely, VUS located close to the head domain had more severe clinical outcomes (Toepfer et al., Reference Toepfer, Garfinkel, Venturini, Wakimoto, Repetti, Alamo, Sharma, Agarwal, Ewoldt, Cloonan, Letendre, Lun, Olivotto, Colan, Ashley, Jacoby, Michels, Redwood, Watkins, Day, Staples, Padrón, Chopra, Ho, Chen, Pereira, Seidman and Seidman2020). It would be interesting to determine if VUS within IDR regions near the head domain had similar impacts on myosin function. In addition, HCM mutations are reported in the S2 motif (R870H, E924K, and E930Δ), which were shown to reduce myosin binding to MyBPC3 (Singh et al., Reference Singh, McNamara and Sadayappan2021). These mutations uncouple the coiled-coil structure upon addition of denaturant as evidenced by CD. These findings raise the possibility that other variants may induce disorder that disrupts myosin function. Adjacent to the S2 motif, the light meromyosin (LMM) region of myosin also harbors disease mutations. Parker et al. (Reference Parker, Batchelor, Wolny, Hughes, Knight and Peckham2018) combined experimental assays and MD simulations to show that two disease mutations, A1603P and K1617Δ, in the LMM motif reduce coiled-coil helicity and lead to abnormal sarcomere assembly. Large-scale gene sequencing has additionally identified mutations in MYH6 and MYH7 that are associated with congenital heart disease (Jin et al., Reference Jin, Homsy, Zaidi, Lu, Morton, DePalma, Zeng, Qi, Chang, Sierant, Hung, Haider, Zhang, Knight, Bjornson, Castaldi, Tikhonoa, Bilguvar, Mane, Sanders, Mital, Russell, Gaynor, Deanfield, Giardini, Porter, Srivastava, Lo, Shen, Watkins, Yandell, Yost, Tristani-Firouzi, Newburger, Roberts, Kim, Zhao, Kaltman, Goldmuntz, Chung, Seidman, Gelb, Seidman, Lifton and Brueckner2017; Theis et al., Reference Theis, Hu, Sundsbak, Evans, Bamlet, Qureshi, O’Leary and Olson2021; Yu et al., Reference Yu, Huang, Zhang, Wang, Chen and Cheng2021).

The myosin heavy chains contain several IDRs, which play important roles in binding to actin (Robert-Paganin et al., Reference Robert-Paganin, Pylypenko, Kikuti, Sweeney and Houdusse2020). These IDRs have been more extensively studied in both cardiac and non-muscle myosin isoforms (Risi et al., Reference Risi, Eisner, Belknap, Heeley, White, Schröder and Galkin2017; Doran et al., Reference Doran, Pavadai, Rynkiewicz, Walklate, Bullitt, Moore, Regnier, Geeves and Lehman2020). Risi et al. constructed a structural model of the cardiac actomyosin complex by fitting available high-resolution myosin/thin-filament structures (von der Ecken et al., Reference von der Ecken, Heissler, Pathan-Chhatbar, Manstein and Raunser2016) to the cryo-EM density of the cardiac thin filament (Risi et al., Reference Risi, Eisner, Belknap, Heeley, White, Schröder and Galkin2017). This model shows key structural motifs of the myosin, such as the highly dynamic loop 4, which has direct contact with the thin filament. More importantly, this cardiac model shows that tropomyosin assumes a different angle compared with that in skeletal model, which may explain the higher activation potential of cardiac filament by Ca2+ (Risi et al., Reference Risi, Eisner, Belknap, Heeley, White, Schröder and Galkin2017). Similarly, a state model of the skeletal myosin/F-actin/tropomyosin complex was built via a combination of docking/cryo-EM fitting and MD simulations (Doran et al., Reference Doran, Pavadai, Rynkiewicz, Walklate, Bullitt, Moore, Regnier, Geeves and Lehman2020). This model provides atomic-level resolution for the myosin motor functional cycle and shows that the interactions between myosin's dynamics loop 4 (amino acids 354–380) and the thin filament are crucial for myosin motor activation (Doran et al., Reference Doran, Pavadai, Rynkiewicz, Walklate, Bullitt, Moore, Regnier, Geeves and Lehman2020).

A complex between myosin VI (MYO6) with F-actin at 4.6 Å via cryo-EM spectroscopy and MD simulations was reported (Gurel et al., Reference Gurel, Kim, Ruijgrok, Omabegho, Bryant and Alushin2017) that could yield mechanistic clues for cardiac myosin. Specifically, the study suggested that two disordered loops form essential interactions with actin that stabilize the complex, but were not resolved (Gurel et al., Reference Gurel, Kim, Ruijgrok, Omabegho, Bryant and Alushin2017). Importantly, one of these loops (T392–P410) is homologous with cardiac myosins and is a locus for several HCM-causing mutations (Gurel et al., Reference Gurel, Kim, Ruijgrok, Omabegho, Bryant and Alushin2017). Based on studies in scallop striated muscle myosin, a helix motif (C693–F707) undergoes a disorder-to-ordered transition during its functional cycle (Houdusse et al., Reference Houdusse, Kalabokis, Himmel, Szent-Gyorgyi and Cohen1999). The human cardiac myosin (MYH7) likely exhibits the same transition given that the motif is conserved. Similarly, the α-helical coiled-coil structure of Myosin Va (MYO5a) only folds upon binding to the myosin Va light chain (Wagner et al., Reference Wagner, Fodor, Ginsburg and Hammer2006), which may also occur in cardiac myosin. Along these lines, MD simulations have shown that the N-terminal IDR of smooth muscle myosin regulatory light chain undergoes a disorder-to-ordered transition upon phosphorylation (Espinoza-Fonseca et al., Reference Espinoza-Fonseca, Kast and Thomas2007), highlighting the importance of PTMs in regulating myosin function.

Atomistic structures of the myosin regulatory light chains (MYL2 and MYL3) in complex with the cardiac myosin heavy chain (MYH7) have been obtained via homology modeling (PDB 5TBY (Alamo et al., Reference Alamo, Ware, Pinto, Gillilan, Seidman, Seidman and Padron2017), Fig. 2). While the MYL2 structure is mostly complete with only a few unresolved N-terminal residues, the MYL3 structrure is missing nearly 40 N-terminal residues (Alamo et al., Reference Alamo, Ware, Pinto, Gillilan, Seidman, Seidman and Padron2017), which suggests that the region is intrinsically disordered. To our knowledge, IDR studies of the myosin light chain isoforms have not been reported.

Cardiac myosin-binding protein C (MyBPC3) is a 137 kD protein that bridges the thin and thick filament (Oakley et al., Reference Oakley, Hambly, Curmi and Brown2004). It is generally thought to simultaneously modulate myosin availability to bind actin as well as the availability of myosin-binding sites on actin (Heling et al., Reference Heling, Geeves and Kad2020). In this capacity, its chief interaction partners are myosin, actin, and titin (Oakley et al., Reference Oakley, Hambly, Curmi and Brown2004). Details continue to emerge, but there is a growing appreciation that MyBPC3 maintains the thick filament ‘off states’ and thin filament ‘on states’ (Kampourakis et al., Reference Kampourakis, Yan, Gautel, Sun and Irving2014) that prevail during diastole and systole, respectively. At low Ca2+, MyBPC3 may also sequester myosin heads in a super-relaxed state, which describes a shift of their conformational ensemble from actin to the myosin tails of the thick filament (Palmer et al., Reference Palmer, Sadayappan, Wang, Weith, Previs, Bekyarova, Irving, Robbins and Maughan2011). Using skinned myocardial strips experiments, Tanner et al. showed that phosphorylation of cardiac MyBPC3 accelerates the rates of myosin detachment from thin filament (Tanner et al., Reference Tanner, Previs, Wang, Robbins and Palmer2021), suggesting MyBPC3 plays a key role in regulating myofilament force generation. The abundance of HCM-causing mutations identified on MyBPC3 (Harris et al., Reference Harris, Lyons and Bezold2011), including 153 pathogenic or likely pathogenic variants located in its predicted IDRs (Fig. 3), have been the topic of studies using both experimental and computational methodologies (reviewed in Sequeira et al., Reference Sequeira, Witjas-Paalberends, Kuster and van der Velden2014; Main et al., Reference Main, Fuller and Baillie2020).

Three MyBPC3 isoforms are found in muscle: fast skeletal, slow skeletal, and cardiac. Although all three isoforms share a common domain organization consisting of seven immunoglobulin I-like (Ig) domains and three fibronectin 3-like domains (Flashman et al., Reference Flashman, Redwood, MoolmanSmook and Watkins2004), the cardiac isoform has several additional motifs that are indispensable for its function (Howarth et al., Reference Howarth, Ramisetti, Nolan, Sadayappan and Rosevear2012). For example, MyBPC3's ~270 residue N-terminus comprising the C0–C1 domains is crucial for heart function, as deletion of these residues resulted in DCM in a mouse model (Lynch et al., Reference Kumar, McNamara, Kuster, Sivaguru, Singh, Previs, Lee, Kuffel, Zilliox, Lin, Ma, Gibson, Blaxall, Nieman, Lorenz, Leichter, Leary, Janssen, Tombe, Gilbert, Craig, Irving, Warshaw and Sadayappan2021). The modular nature of MyBPC3 likely has the advantage that it allows the protein's function to be tuned by modifying its domain properties. Napierski et al. for instance demonstrate that an MyBPC3 construct lacking the C0–C7 domains exhibits abnormal function, which is rescued by inserting recombinant C0–C7 domains (Napierski et al., Reference Napierski, Granger, Langlais, Moran, Strom, Touma and Harris2020). Another unique motif in cardiac MyBPC3 is the ~100-amino acid M-domain that bridges its C1 and C2 domains (Howarth et al., Reference Howarth, Ramisetti, Nolan, Sadayappan and Rosevear2012) and binds myosin (Singh et al., Reference Singh, McNamara and Sadayappan2021). The N-terminal fragment of the M-domain was identified as an IDR by CD and NMR (Howarth et al., Reference Howarth, Ramisetti, Nolan, Sadayappan and Rosevear2012). Atomic force microscopy (AFM) studies additionally indicate that its phosphorylation reduces the M-domain's extensibility and likely attenuates MyBPC3's ability to regulate cardiac muscle contraction (Previs et al., Reference Previs, Mun, Michalek, Previs, Gulick, Robbins, Warshaw and Craig2016). Intriguingly, NMR studies have also identified a highly flexible linker in the M-domain that serves as a major binding site for the regulatory calmodulin (CaM) (Michie et al., Reference Michie, Kwan, Tung, Guss and Trewhella2016), which we speculate may confer Ca2+-dependent conformational changes to its IDR. Additionally, Colson et al. (Reference Colson, Thompson, Espinoza-Fonseca and Thomas2016) used MD simulations to demonstrate that M-domain (residues T255–R357) phosphorylation reduced its structural disorder, stabilized the C0/C1 domain folds, and exposed a cryptic protein–protein binding site. This finding provides mechanistic insight into how regulatory control via the PTMs within IDRs in the M-domain impacts cardiac contractility.

In addition to the M-domain, other IDRs in MyBPC3 are speculated to exist. AFM analyses from Karsai et al. indicate that the force–extension relationship of the intact MyBPC3 is heterogeneous, with some regions extending more easily than would be expected for folded Ig domains (Karsai et al., Reference Karsai, Kellermayer and Harris2011). These regions are believed to be disordered and likely correspond to the linkers connecting MyBPC3's folded domains. Doh et al. (Reference Doh, Bharambe, Holmes, Dominic, Swanberg, Mamidi, Chen, Bandyopadhyay, Ramachandran and Stelzer2022) showed that the flexible linker connecting the C4 and C5 domains of MyBPC3 not only modulates the secondary structure content in the C4 and C5 domains, but also affects their relative interdomain orientations and associated kinetics (Doh et al., Reference Doh, Bharambe, Holmes, Dominic, Swanberg, Mamidi, Chen, Bandyopadhyay, Ramachandran and Stelzer2022). By using MD simulations, they pinpointed specific residue–residue contacts, and the local conformation changes of the linker, that contribute to its modulatory role. Another study focusing on the interdomain flexible linker from Potrzebowski et al. (Reference Potrzebowski, Trewhella and Andre2018) combined Bayes inference and molecular simulations to build structural models of an MyBPC3 construct including an M-domain fragment, C2-domain, and a linker. The structural models that best match the experimental SAXS data show diverse interdomain orientations, demonstrating the flexible linker's role in supporting a broad conformation ensemble. Interestingly, genome scale bioinformatic analysis has also revealed that MyBPC3 splice isoforms tend to overlap with disordered regions (Lau et al., Reference Lau, Han, Williams, Thomas, Shrestha, Wu and Lam2019), which implicates the IDRs in myofilament function.

Titin (TTN) is a behemoth protein (ca. 3816 kD) that secures the thick filament to the Z-disk, by spanning one-half sarcomere to the M-line (LeWinter and Granzier, Reference LeWinter and Granzier2010). Titin is highly modular in that it contains folded Ig domains interspersed with unstructured regions such as PEVK motifs (Linke et al., Reference Linke, Ivemeyer, Mundel, Stockmeier and Kolmerer1998). The PEVK repeats are intrinsically disordered, ~28 residue motifs enriched in proline, glutamic acid, valine, and lysine (Ma and Wang, Reference Ma and Wang2003). These IDRs contribute to titin's passive elasticity (Linke et al., Reference Linke, Kulke, Li, Fujita-Becker, Neagoe, Manstein, Gautel and Fernandez2002; Ma and Wang, Reference Ma and Wang2003) and are substrates for proteins like S100A1 (Yamasaki et al., Reference Yamasaki, Berri, Wu, Trombitas, McNabb, Kellermayer, Witt, Labeit, Labeit, Greaser and Granzier2001). It has been speculated that S100A1 binding at these repeats modulates passive tension (Granzier et al., Reference Granzier, Fukushima and Chung2010). The PEVK repeat was observed to have mostly disordered secondary structure (Poly II helix, b-turn and coils) and larger Stokes radius using CD, gel permeation chromatography, and gel electrophoresis, which confirm its intrinsically disordered nature (Ma and Wang, Reference Ma and Wang2003; Duan et al., Reference Duan, DeKeyser, Damodaran and Greaser2006). Interestingly, idiopathic restrictive cardiomyopathy mutations have been found in the PEVK motifs as well as the fibronectin-type III (FnIII) domains, of which the latter are also likely to be disordered (Tarnovskaya et al., Reference Tarnovskaya, Kiselev, Kostareva and Frishman2017). Mutations in these regions were predicted via PONDR-FIT (Xue et al., Reference Xue, Dunbrack, Williams, Dunker and Uversky2010) to alter their disorder (Tarnovskaya et al., Reference Tarnovskaya, Kiselev, Kostareva and Frishman2017).

Beyond titin's PEVK motifs, linkers connecting its modular domains are likely IDRs based on predictions from PONDR (Fig. 3). Simulations from the Schulten lab investigated the effect of the PEVK domains and domain unfolding on tension, which led to TTN's characterization as an entropic spring (Lu et al., Reference Lu, Isralewitz, Krammer, Vogel and Schulten1998; Lee et al., Reference Lee, Hsin, Mayans and Schulten2007). In addition, steered MD simulations have been extensively performed on TTN constructs consisting of multiple Ig domains and interdomain linkers (reviewed in Hsin et al., Reference Hsin, Strümpfer, Lee and Schulten2011). These simulations put forth a structural basis for TTN's impressive plasticity, how interdomain bending is mostly mediated by flexible linkers, and how domain unfolding may influence plasticity. Moreover, IDRs in TTN could also serve as binding motifs for protein–protein interactions (PPIs). As an example, the N2A titin isoform contains long IDR linkers flanking its binding site for ankyrin repeat proteins that were determined to be intrinsically disordered using NMR and HDXMS (Zhou et al., Reference Zhou, Fleming, Lange, Hessel, Bogomolovas, Stronczek, Grundei, Ghassemian, Biju, Borgeson, Bullard, Linke, Chen, Kovermann and Mayans2021a). ClinVar and PhosphoSitePlus identify more than 400 pathogenic/likely pathogenic variants and over 300 PTMs within its predicted IDRs (Fig. 3). This concurs with suggestions that its defects are responsible for the majority of dilated cardiomyopathies (Herman et al., Reference Herman, Lam, Taylor, Wang, Teekakirikul, Christodoulou, Conner, DePalma, McDonough, Sparks, Teodorescu, Cirino, Banner, Pennell, Graw, Merlo, Di Lenarda, Sinagra, Bos, Ackerman, Mitchell, Murry, Lakdawala, Ho, Barton, Cook, Mestroni, Seidman and Seidman2012).

MAPIDs of the Z-disk

The Z-disk is a central hub (Sols and Solaro, Reference Sols and Solaro2021) that interfaces adjacent sarcomeres and sarcomeres to organelle membranes (see Fig. 1a). It is a nexus for sensing changes in mechanical demand and can communicate these changes to myriad signaling pathways to regulate cardiac function (Sols and Solaro, Reference Sols and Solaro2021). For this reason, it is also a prime target for a number of regulatory mechanisms (Sols and Solaro, Reference Sols and Solaro2021). Unlike the thin and thick filament for which the primary constituents have largely been identified, proteins composing and interacting with the Z-disk are continuing to be found.

α-Actinin (ACTN2) is a large (104 kD), dimeric protein that interfaces with titin and actin filaments, where it contributes to sarcomere assembly (Chopra et al., Reference Chopra, Kutys, Zhang, Polacheck, Sheng, Luu, Eyckmans, Hinson, Seidman, Seidman and Chen2018). α-actinin cross-links actin filaments with the Z-disk (Maruyama and Ebashi, Reference Maruyama and Ebashi1965) and also competes with calsarcin to bind calcineurin (CN) (Frey et al., Reference Frey, Richardson and Olson2000; Seto et al., Reference Seto, Quinlan, Lek, Zheng, Garton, MacArthur, Hogarth, Houweling, Gregorevic, Turner, Cooney, Yang and North2013). α-actinin's structure is almost entirely resolved at 3.5 Å resolution for residues Y19–L894 (PDB 4D1E (Ribeiro et al., Reference Ribeiro, Pinotsis, Ghisleni, Salmazo, Konarev, Kostan, Sjöblom, Schreiner, Polyansky, Gkougkoulia, Holt, Aachmann, Zagrović, Bordignon, Pirker, Svergun, Gautel and Djinović-Carugo2014)). In its dimeric state, the structure does not present disordered structural domains. However, PONDR suggests α-actinin contains several IDRs, which perhaps may be evident in the isolated monomeric structure. Thirty-four PTMs have been reported by the PhosphoSitePlus, 17 of which are located in the predicted IDRs (Fig. 3). Fourteen pathogenic or likely pathogenic variants are reported in ClinVar, with five in the predicted IDRs.

Crystallin alpha B (CRYAB) CRYAB is a 20 kD, ubiquitously-expressed, small heat shock protein that regulates cellular responses to stress (Dimauro et al., Reference Dimauro, Antonioni, Mercatelli and Caporossi2018). As a chaperone, CRYAB associates with misfolded proteins to suppress their aggregation (Dimauro et al., Reference Dimauro, Antonioni, Mercatelli and Caporossi2018). At least under ischemic conditions, it binds with titin to potentially protect the protein from degradation (Golenhofen et al., Reference Golenhofen, Arbeiter, Koob and Drenckhahn2002). The protein consists of three domains, the N-terminal domain (NTD, residues M1–S59), the α-crystallin domain (ACD, residues W60–K150), and the C-terminal domain (CTD, residues Q151–K175) (Braun et al., Reference Braun, Zacharias, Peschek, Kastenmuller, Zou, Hanzlik, Haslbeck, Rappsilber, Buchner and Weinkauf2011). Although its complete structure is deposited at 9.40 Å resolution from cryo-EM as 24-meric alphaB-crystallin (Braun et al., Reference Braun, Zacharias, Peschek, Kastenmuller, Zou, Hanzlik, Haslbeck, Rappsilber, Buchner and Weinkauf2011), higher-resolution NMR structures reveal only its ACD domain, while the NTD and CTD remain unresolved (Jehle et al., Reference Jehle, Rajagopal, Bardiaux, Markovic, Kühne, Stout, Higman, Klevit, van Rossum and Oschkinat2010). NMR experiments indicate that its C-terminus is intrinsically disordered and may self-aggregate (Baldwin et al., Reference Baldwin, Walsh, Hansen, Hilton, Benesch, Sharpe and Kay2012); this aggregation is considered a pathological marker in histologies (Zhang et al., Reference Zhang, Rajasekaran, Orosz, Xiao, Rechsteiner and Benjamin2010). Three PTM sites S19, S45, and S59 have been identified in its N-domain (Chiappori et al., Reference Chiappori, Mattiazzi, Milanesi and Merelli2016). MD simulations reveal that phosphorylation of S59 alters a key interface for its multimeric self-assembly (Chiappori et al., Reference Chiappori, Mattiazzi, Milanesi and Merelli2016). Interestingly, three cataract-associated mutations were also found in this region, which were shown to both alter self-assembly and its interaction with other proteins (Muranova et al., Reference Muranova, Strelkov and Gusev2020). In addition, nearly 10 other muscle disease-associated mutations have been reported in its ACD and CTD (Dimauro et al., Reference Dimauro, Antonioni, Mercatelli and Caporossi2018).

Enigma homolog isoform 2 (ENH2) is a 64 kD protein that interacts with calsarcin in the Z-disk and α-actinin (Cheng et al., Reference Cheng, Kimura, Peter, Cui, Ouyang, Shen, Liu, Gu, Dalton, Evans, Knowlton, Peterson and Chen2010). As a splice isoform of the PDLIM5 gene (Huang et al., Reference Huang, Qu, Ouyang, Zhong and Dai2020a), ENH2 contains a postsynaptic density protein of 95 (PSD-95) PSD-95/Discs large/Zonula occludens-1 domains (PDZ) domain and three LIM domains. Of these, the PDZ domain plays important roles in signal transduction through PPIs formed with targets (Ivarsson, Reference Ivarsson2012). The LIM domain has two zinc fingers and also engages a diverse range of signaling pathways. The PDZ and LIM domains are generally folded (Elkins et al., Reference Elkins, Gileadi, Shrestha, Phillips, Wang, Muniz and Doyle2010), which is consistent with our PONDR predictions of ENH2 in Fig. 3. The regions between the PDZ and LIM domains are predicted to be IDRs, which agrees with annotations from UniProt for EHN2. Emerging evidence also suggests that even the folded LIM domain likely contains IDRs, as an unpublished solution NMR structure of the LIM domain (PDB 2DAR) exhibits highly dynamic N- and C-termini.

Knock-out of the enigma homolog protein leads to impaired cardiac contraction and DCM in a mouse model (Cheng et al., Reference Cheng, Kimura, Peter, Cui, Ouyang, Shen, Liu, Gu, Dalton, Evans, Knowlton, Peterson and Chen2010). Top–down proteomics, which utilizes mass spectrometry to characterize intact proteins, demonstrate that changes in ENH2 phosphorylation at S118 occur in ischemia (Peng et al., Reference Peng, Gregorich, Valeja, Zhang, Cai, Chen, Guner, Chen, Schwahn, Hacker, Liu and Ge2014) and HCM (Tucholski et al., Reference Tucholski, Cai, Gregorich, Bayne, Mitchell, McIlwain, Lange, Wrobbel, Karp, Hite, Vikhorev, Marston, Lal, Li, Remedios, Kohmoto, Hermsen, Ralphe, Kamp, Moss and Ge2020). PhosphoSitePlus suggests 32 PTM sites in the predicted IDR regions, while ClinVar does not report any pathogenic or likely pathogenic variants that change the protein sequence.

Obscurin (OBSCN) is an 869 kD protein that links myofibrils to the SR (Lange et al., Reference Lange, Ouyang, Meyer, Cui, Cheng, Lieber and Chen2009) and possibly to the cytoskeleton (Geisler et al., Reference Geisler, Robinson, Hauringa, Raeker, Borisov, Westfall and Russell2007). Similar to other modular proteins like titin and nebulin, obscurin consists of many folded domains joined by disordered linkers (Young et al., Reference Young, Ehler and Gautel2001). MD simulations have shown that the linkers between the modular domains of obscurin are flexible, which allows the bridged folded domains to sample broad conformational ensembles that likely contribute to the protein's elasticity (Whitley et al., Reference Whitley, Ex-Willey, Marzolf, Ackermann, Tongen, Kokhan and Wright2019). In addition, a solution NMR structure of obscurin's PDZ domain (residues R3614–P3713, PDB 2EDH) is highly dynamic, which is consistent with PONDR predictions in Fig. 3. Of the putative PTMs found in PhosphoSitePlus (see Fig. 3), 54 are found within these IDRs. While ClinVar reports just one pathogenic or likely pathogenic variant that changes the sequence of the predicted IDRs, 16 cardiomyopathy-linked variants spanning the OBSCN sequence were reported in 2017 (Marston, Reference Marston2017). Given the many intermittent IDR regions predicted in OBSCN (Fig. 3), these variants are likely to be located within or immediately adjacent to IDR regions. As OBSCN's role in the sarcomere continues to be clarified, new cardiomyopathy-linked variants may continue to emerge (Marston, Reference Marston2017).

Myotilin (MYOT) is a pseudonym for limb-girdle muscular dystrophy 1A (LGMD1A) protein (Salmikangas et al., Reference Salmikangas, van der Ven, Lalowski, Taivainen, Zhao, Suila, Schröder, Lappalainen, Fürst and Carpén2003). The 55 kD protein localizes to the Z-disk and cross-links α-actinin, where it is believed to contribute to the assembly of actin filaments (Salmikangas et al., Reference Salmikangas, van der Ven, Lalowski, Taivainen, Zhao, Suila, Schröder, Lappalainen, Fürst and Carpén2003). The N-terminal fragment and C-tail of myotilin are predicted to be disordered and contain binding sites for proteins such as α-actinin-2 and Z-disk-associated, alternatively spliced, PDZ motif-containing protein (ZASP) (Puž et al., Reference Puž, Pavšič, Lenarčič and Djinović-Carugo2017). This prediction is in excellent agreement with our PONDR scores for MYOT in Fig. 3. The presence of IDRs in MYOT is further confirmed by an unpublished solution NMR structure of myotilin's C-terminal fragment (PDB 2KKQ) that shows highly dynamic regions. The IDR in the N-terminal fragment of MYOT harbors several muscle disorder associated mutations (Puž et al., Reference Puž, Pavšič, Lenarčič and Djinović-Carugo2017), and PhosphoSitePlus reveals five PTMs in its predicted IDRs.

Myomesin 1 (MYOM1) is a 188 kD protein that together with titin and obscurin form the dynamic ‘M’ band found between the Z-disks of the cardiac sarcomere (Lamber et al., Reference Lamber, Guicheney and Pinotsis2022). In this arrangement, the protein plays an important role in sarcomere organization (Lamber et al., Reference Lamber, Guicheney and Pinotsis2022). Myomesin 1 is also suggested to bind myosin (Lamber et al., Reference Lamber, Guicheney and Pinotsis2022), which could directly impact sarcomere contraction. Predictions using IUPred2A (Mészáros et al., Reference Mészáros, Erdos and Dosztányi2018) reveal that MYOM contains IDRs, while genome scale bioinformatic analysis suggests MYOM splice isoforms frequently overlap with IDRs (Lau et al., Reference Lau, Han, Williams, Thomas, Shrestha, Wu and Lam2019). One of these splice isoforms contains an insertion of an ~100 residue elastic segment (EH domain) at the center of the protein. This EH-myosin is the main component of the M-band in higher vertebrates (Schoenauer et al., Reference Schoenauer, Emmert, Felley, Ehler, Brokopp, Weber, Nemir, Faggian, Pedrazzini, Falk, Hoerstrup and Agarkova2011). Expression of the EH-myomesin is significantly up-regulated in DCM patients (Schoenauer et al., Reference Schoenauer, Emmert, Felley, Ehler, Brokopp, Weber, Nemir, Faggian, Pedrazzini, Falk, Hoerstrup and Agarkova2011). Moreover, AFM, transmission electron microscopy, and CD data suggest that the EH segment is disordered and contributes to the protein's elasticity (Schoenauer et al., Reference Schoenauer, Bertoncini, Machaidze, Aebi, Perriard, Hegner and Agarkova2005). PONDR scores for MYOM1 in Fig. 3 reveal multiple IDRs spanning the protein. While PhosphoSitePlus reports 15 PTMs in the predicted IDRs, no variants that alter its IDR sequences are reported in Clinvar.

Desmin (DES) is a 54 kD protein that anchors the myofibrils by interconnecting the Z-disks to the cell cytoskeleton (Goldfarb et al., Reference Goldfarb, Park, Cervenáková, Gorokhova, Lee, Vasconcelos, Nagle, Semino-Mora, Sivakumar and Dalakas1998; Brodehl et al., Reference Brodehl, Dieding, Klauke, Dec, Madaan, Huang, Gargus, Fatima, Saric, Cakar, Walhorn, Tonsing, Skrzipczyk, Cebulla, Gerdes, Schulz, Gummert, Svendsen, Olesen, Anselmetti, Christensen, Kimonis and Milting2013). It serves both signaling and structural roles in cardiomyocytes (McLendon and Robbins, Reference McLendon and Robbins2011). Mutations in desmin are associated with desmin-related (cardio)myopathy, which is also known as desminopathy (Goldfarb et al., Reference Goldfarb, Olivé, Vicart, Goebel and Laing2008). These defects are accompanied by aggregates of misfolded proteins (McLendon and Robbins, Reference McLendon and Robbins2011), which suggests potential correlations between desmin structural defects and protein quality control. Predictions using DISOPRED3 and DICHOT reveal that both the N- and C-terminus of desmin are intrinsically disordered (Anbo et al., Reference Anbo, Sato, Okoshi and Fukuchi2019). This prediction is in agreement with our PONDR scores (Fig. 3). Moreover, the C-tail IDR harbors the binding site for the chaperone alpha B-crystallin (CRYAB) (Anbo et al., Reference Anbo, Sato, Okoshi and Fukuchi2019). PhosphoSitePlus and ClinVar report 22 PTMs and 47 pathogenic or likely pathogenic variants, respectively, in the predicted IDRs.

Four and a half LIM domains 2 (FHL2) is a 32 kD protein that serves as a biochemical stress sensor in the sarcomere, in part through its interactions with titin (Sheikh et al., Reference Sheikh, Raskin, Chu, Lange, Domenighetti, Zheng, Liang, Zhang, Yajima, Gu, Dalton, Mahata, Brown, Peterson, Omens, McCulloch and Chen2008). In principle it may also regulate phosphorylase activity given its binding interactions with CN phosphatase (Hojayev et al., Reference Hojayev, Rothermel, Gillette and Hill2012). According to several studies (Chu et al., Reference Chu, Bardwell, Gu, Ross and Chen2000; Okamoto et al., Reference Okamoto, Li, Noma, Hiroi, Liu, Taniguchi, Ito and Liao2013; Friedrich et al., Reference Friedrich, Reischmann, Schwalm, Unger, Ramanujam, Münch, Müller, Hengstenberg, Galve, Charron, Linke, Engelhardt, Patten, Richard, van der Velden, Eschenhagen, Isnard and Carrier2014), the protein is believed to play a minimal role during normal cardiac development, but may limit the development of cardiac hypertrophy. Our PONDR scores for FHL2 show negligible IDR content, which is consistent with annotations from UniProtKB. However, solution NMR structures of its LIM domains are highly dynamic at their N/C-termini (PDB 2D8Z, 1X4L, 1X4K) and therefore may be characterized as IDRs. While no pathogenic/likely pathogenic variants that change the IDR sequence have been reported in the ClinVar database, PhosphoSitePlus reports five PTMs within the LIM termini.

Nebulin (NEB) and nebulette (NEBL). Nebulette belongs to the nebulin family, but has a much smaller size of 116 kD compared with the 773 kD nebulin (Bang and Chen, Reference Bang and Chen2015). Nebulin is primarily expressed in skeletal muscle while nebulette is highly expressed in cardiac muscle (Bang and Chen, Reference Bang and Chen2015). Nebulin and nebulette are modular proteins that share a common building block called the ‘nebulin repeat module’ (Bang and Chen, Reference Bang and Chen2015). Sequence comparisons between nebulette and nebulin from different species show that their basic building blocks are highly conserved (Moncman and Wang, Reference Moncman and Wang2000). Similar to nebulin (Labeit and Kolmerer, Reference Labeit and Kolmerer1995), cardiac nebulette bridges actin to desmin (Hernandez et al., Reference Hernandez, Bennett, Dunina-Barkovskaya, Wedig, Capetanaki, Herrmann and Conover2016) and regulates Z-disk assembly (Moncman and Wang, Reference Moncman and Wang1999). Studies of IDRs in nebulette have not yet been reported, therefore we refer to work focused on its homolog, nebulin.

Nebulin is a ~1 μm long intrinsically disordered scaffolding protein along the thin filaments, with a significant percentage of its residues predicted as IDRs by the PONDR-VL3H algorithm (Wu et al., Reference Wu, Forbes and Wang2016). Its IDRs are suggested to regulate sarcomere assembly, given the close alignment between the periodicity of high disorder scores with the positions of myosin-associated proteins in the A-band (Wu et al., Reference Wu, Forbes and Wang2016). These disordered regions most likely reside between the nebulin building blocks. PTM sites have been identified in both nebulin and nebulette throughout their sequences, while most of those sites tend to localize to regions that share high homology (Moncman and Wang, Reference Moncman and Wang2000). Missense mutations of nebulin are mainly related to nemaline myopathy, while several mutations in nebulette are linked to familial and idiopathic dilated cardiomyopathy (Bang and Chen, Reference Bang and Chen2015). Given the many intermittent IDRs in nebulin predicted by us (Fig. 3) and by Wu et al. (Reference Wu, Forbes and Wang2016), it is likely that some of these disease-associated mutations fall within its IDRs. ClinVar reports 61 pathogenic/likely pathogenic variants in the predicted IDRs, in addition to 37 PTMs reported in the PhosphoSitePlus database (Fig. 3).

Myopalladin (MYPN) is a 145 kD protein that bridges nebulin (skeletal) or nebulette (cardiac) to α-actinin in the Z-disk (Bang et al., Reference Bang, Mudry, McElhinny, Trombitas, Geach, Yamasaki, Sorimachi, Granzier, Gregorio and Labeit2001). Myopalladin is a modular protein that contains five Ig-like domains (Bang et al., Reference Bang, Mudry, McElhinny, Trombitas, Geach, Yamasaki, Sorimachi, Granzier, Gregorio and Labeit2001); linkers connecting the modular Ig-like domains are likely to be IDRs based on the PONDR data in Fig. 3. Studies targeting these potential IDRs have not been reported in the literature. However, genetic screens of patients with cardiomyopathy have identified dozens of mutations in MYPN (Duboscq-Bidot et al., Reference Duboscq-Bidot, Xu, Charron, Neyroud, Dilanian, Millaire, Bors, Komajda and Villard2007; Purevjav et al., Reference Purevjav, Arimura, Augustin, Huby, Takagi, Nunoda, Kearney, Taylor, Terasaki, Bos, Ommen, Shibata, Takahashi, Itoh-Satoh, McKenna, Murphy, Labeit, Yamanaka, Machida, Park, Alexander, Weintraub, Kitaura, Ackerman, Kimura and Towbin2012), of which two fall within its IDRs. Several of these mutations cause disrupted Z-disk assembly along with abnormal co-expression of MYPN with α-actinin, desmin, and ankyrin (Purevjav et al., Reference Purevjav, Arimura, Augustin, Huby, Takagi, Nunoda, Kearney, Taylor, Terasaki, Bos, Ommen, Shibata, Takahashi, Itoh-Satoh, McKenna, Murphy, Labeit, Yamanaka, Machida, Park, Alexander, Weintraub, Kitaura, Ackerman, Kimura and Towbin2012). The PhosphoSitePlus reports 35 PTMs in the predicted IDRs.

F-actin-capping protein subunit alpha-2 (CAPZA2) is a 33 kD protein localized to the Z-disk that directs the orientation and direction of actin during muscle fiber development (Solis and Russell, Reference Solis and Russell2019). It does so by regulating the growth of the barbed end of actin filaments (Funk et al., Reference Funk, Merino, Schaks, Rottner, Raunser and Bieling2021). To date, variants in CAPZA2 have not been linked to cardiac disorders, although there have been associations established with mental impairment (Huang et al., Reference Huang, Mao, Jaarsveld, Shu, Terhal, Jia, Xi, Peng, Yan, Yuan, Li, Wang and Bellen2020b). IDR-related studies have not been reported for CAPZA2, although based on Fig. 3, there is evidence of IDRs within residues M1–F20, D99–V124, and R259–W271. ClinVar and PhosphoSitePlus report two pathogenic/likely pathogenic variants (K256E and R259L) and one PTM (S4) located near the predicted IDRs.

Miscellaneous MAPIDs of the sarcomere

Several proteins are associated with the myofilaments but are not unambiguously classified into the Z-disk, thick filament, or thin filament assemblies. Many contribute to sarcomere assembly, for instance through cross-linking proteins, or play important signaling roles, such as mechanosensing.

B type ankyrin (ANK2) is a 434 kD protein that bridges the sarcomere M-line to the cell membrane via obscurin (Cunha and Mohler, Reference Cunha and Mohler2008), where it interacts with ion channels and transporters. Its dysfunction is linked to a variety of electrical defects in cardiac function (Sucharski et al., Reference Sucharski, Dudley, Keith, El Refaey, Koenig and Mohler2020). Its N-terminus contains well-folded 24 ankyrin repeats, which form a groove that can bind many IDR-containing membrane proteins (Wang et al., Reference Wang, Wei, Chen, Ye, Yu, Bennett and Zhang2014). Its C-terminus is intrinsically disordered as evidenced by little to no secondary structure content in CD studies and its large Stokes radius (Abdi et al., Reference Abdi, Mohler, Davis and Bennett2006). The ankyrin repeats are auto-inhibited by several IDR fragments from its C-terminus, which was determined via binding assays and X-ray crystallography (Chen et al., Reference Chen, Li, Wang, Wei and Zhang2017). Two idiopathic restrictive cardiomyopathy mutations are found in ankyrin's IDRs (Tarnovskaya et al., Reference Tarnovskaya, Kiselev, Kostareva and Frishman2017) in addition to about 80 PTM sites via PhosphoSitePlus.

Cysteine-rich protein 3 (CRIP3) is a 24 kD protein and together with CRIP1 and CRIP2, it forms the cysteine-rich intestinal protein (Crip) family (Hempel and Kuhl, Reference Hempel and Kuhl2014). CRIP3 and its homolog CRIP2 are expressed in cardiovascular tissue (Wei et al., Reference Wei, Lin, Lu, Chen and You2011; Hempel and Kuhl, Reference Hempel and Kuhl2014), where they are speculated to have a role in mechanosensing (Boateng et al., Reference Boateng, Belin, Geenen, Margulies, Martin, Hoshijima, de Tombe and Russell2007). Its precise localization in the sarcomere is not established, although the STRING database of PPIs (Szklarczyk et al., Reference Szklarczyk, Gable, Lyon, Junge, Wyder, Huerta-Cepas, Simonovic, Doncheva, Morris, Bork, Jensen and Mering2019) suggests it may associate with CAPZA3. CRIP3 is composed of two LIM domains connected by a ~30 residue flexible linker. This linker is predicted to be an IDR by our PONDR scores (see Fig. 3). Mutations have been identified in the cysteine-rich protein family that are linked to dilated cardiomyopathy (Knöll et al., Reference Knöll, Hoshijima, Hoffman, Person, Lorenzen-Schmidt, Bang, Hayashi, Shiga, Yasukawa, Schaper, McKenna, Yokoyama, Schork, Omens, McCulloch, Kimura, Gregorio, Poller, Schaper, Schultheiss and Chien2002). Two PTMs, Y132 and S139, located in the second LIM domain are reported in the PhosphoSitePlus database. No pathogenic or likely pathogenic variants in CRIP3 have been reported in ClinVar.

Filamin-C (FLNC) is a 291 kD protein that binds nebulette and ankyrin (Holmes and Moncman, Reference Holmes and Moncman2008; Maiweilidan et al., Reference Maiweilidan, Klauza and Kordeli2011), where it contributes to sarcomere assembly (Agarwal et al., Reference Agarwal, Paulo, Toepfer, Ewoldt, Sundaram, Chopra, Zhang, Gorham, DePalma, Chen, Gygi, Seidman and Seidman2021). Filamin-C contains 24 folded repeats (Ig-like domains) in its C-terminus (Nakamura et al., Reference Nakamura, Stossel and Hartwig2011). These 24 repeats form two rod-like structures, with the first rod assuming an extended chain configuration and the second, a relatively compact folded configuration (Nakamura et al., Reference Nakamura, Stossel and Hartwig2011). These two rod configurations are necessary for filamin's dimerization and binding to actin (Nakamura et al., Reference Nakamura, Stossel and Hartwig2011). Although IDR studies have not yet been reported for filamin-C, an unpublished solution NMR structure of its modular domain (Fig. 4, PDB 2D7O) shows highly dynamic N- and C-termini. PONDR scores for FLNC in Fig. 3 predict that its IDRs are interspersed in the protein's sequence. Based on these predictions, 41 pathogenic or likely pathogenic variants and 36 PTMs are localized to the IDRs, based on the ClinVar and PhosphoSitePlus databases, respectively.

Fig. 4. Solution NMR structures of the common modular domains that are used as building blocks of Z-disk proteins. These structures show highly dynamic termini and evidence of intrinsic disorder in the Z-disk. The PDB IDs are given in parentheses.

Myozenin 2 (MYOZ2 or FATZ-2) is a 39 kD Z-disk protein that binds to and inhibits CN, hence its pseudonym calsarcin (Frey et al., Reference Frey, Richardson and Olson2000; Ruggiero et al., Reference Ruggiero, Chen, Lombardi, Rodriguez and Marian2012). While activation of CN in the adult heart is typically associated with cardiac hypertrophy (Ruggiero et al., Reference Ruggiero, Chen, Lombardi, Rodriguez and Marian2012), mutations in myozenin 2 that are linked to cardiomyopathy and disorganization of the Z-disk may arise independently of CN activity (Ruggiero et al., Reference Ruggiero, Chen, Lombardi, Rodriguez and Marian2012). Sponga et al. found that the MYOZ1 isoform, which shares around 36% sequence identity with MYOZ2, is intrinsically disordered (Sponga et al., Reference Sponga, Arolas, Schwarz, Jeffries, Chamorro, Kostan, Ghisleni, Drepper, Polyansky, Ribeiro, Pedron, Zawadzka-Kazimierczuk, Mlynek, Peterbauer, Doto, Schreiner, Hollerl, Mateos, Geist, Faulkner, Kozminski, Svergun, Warscheid, Zagrovic, Gautel, Konrat and Djinovi-Carugo2021). This was rationalized by the MYOZ1 constructs having higher apparent molecular weights and high percentages of random coil conformations, as measured by size exclusion chromatography and CD, respectively (Sponga et al., Reference Sponga, Arolas, Schwarz, Jeffries, Chamorro, Kostan, Ghisleni, Drepper, Polyansky, Ribeiro, Pedron, Zawadzka-Kazimierczuk, Mlynek, Peterbauer, Doto, Schreiner, Hollerl, Mateos, Geist, Faulkner, Kozminski, Svergun, Warscheid, Zagrovic, Gautel, Konrat and Djinovi-Carugo2021). They additionally obtained a diverse ensemble of atomistic-resolution structures by fitting a pool of randomly generated models to SAXS data (Sponga et al., Reference Sponga, Arolas, Schwarz, Jeffries, Chamorro, Kostan, Ghisleni, Drepper, Polyansky, Ribeiro, Pedron, Zawadzka-Kazimierczuk, Mlynek, Peterbauer, Doto, Schreiner, Hollerl, Mateos, Geist, Faulkner, Kozminski, Svergun, Warscheid, Zagrovic, Gautel, Konrat and Djinovi-Carugo2021). Lastly, using binding assays, NMR, X-ray crystallography, and SAXS, the authors revealed that the intrinsically disordered MYOZ1 forms tight, fuzzy interactions with α-actinin (Sponga et al., Reference Sponga, Arolas, Schwarz, Jeffries, Chamorro, Kostan, Ghisleni, Drepper, Polyansky, Ribeiro, Pedron, Zawadzka-Kazimierczuk, Mlynek, Peterbauer, Doto, Schreiner, Hollerl, Mateos, Geist, Faulkner, Kozminski, Svergun, Warscheid, Zagrovic, Gautel, Konrat and Djinovi-Carugo2021). This raises the possibility that myozenin 2 may exhibit similar interactions with cardiac α-actinin that are important to its function. Our PONDR scores for MYOZ2 indicate that the N-terminal fragment and C-terminus of the protein are IDRs (Fig. 3). Although no pathogenic or likely pathogenic variants were identified in the predicted IDRs of MYOZ2 from the ClinVar database, two HCM mutations, S48P and I246M, have been identified in the N- and C-termini, respectively (Ruggiero et al., Reference Ruggiero, Chen, Lombardi, Rodriguez and Marian2012). The PhosphoSitePlus database reports five PTMs in the predicted IDRs (Fig. 3).

Spectrin beta, erythrocytic (SPTB) is a 247 kD protein that binds to the actin filament (An et al., Reference An, Debnath, Guo, Liu, Lux, Baines, Gratzer and Mohandas2005) and is a major structural component of the cytoskeleton (Winkelmann et al., Reference Winkelmann, Chang, Tse, Scarpa, Marchesi and Forget1990). SPTB binds to α-spectrin to form hetero-tetramers (Long et al., Reference Long, McElheny, Jiang, Park, Caffrey and Fung2007). Although both α and β spectrins are largely characterized as coiled-coil structures (Park et al., Reference Park, Caffrey, Johnson and Fung2003), at least the N-terminus of α-spectrin and residues Q1898–E2083 of β-spectrin were determined to contain IDRs, as determined by CD and NMR studies (Park et al., Reference Park, Caffrey, Johnson and Fung2003; Long et al., Reference Long, McElheny, Jiang, Park, Caffrey and Fung2007). Spectrin isoforms bind at these IDRs and gain helical character after binding (Long et al., Reference Long, McElheny, Jiang, Park, Caffrey and Fung2007). Our PONDR scores suggest that many IDRs span the SPTB sequence, including those already confirmed to be intrinsically disordered (Long et al., Reference Long, McElheny, Jiang, Park, Caffrey and Fung2007). Within these predicted IDRs, PhosphoSitePlus and ClinVar report 13 PTMs and 18 pathogenic or likely pathogenic variants, respectively.

Properties of MAPIDs and their characterization

Advantages and vulnerabilities of IDRs in MAPIDs

Nearly all the proteins we describe in section ‘Myofilament-associated protein with intrinsic disorder (MAPID)s’ have regions that were either determined by experiment to be IDRs or were suggested to be disordered by PONDR. The abundance of these regions suggests that they perform distinct functional roles relative to their folded counterparts. We speculate there are several advantages to these IDRs in myofilament proteins.

Native structure

IDRs' roles in cellular signaling and transcription are well-known (Gibbs and Showalter, Reference Gibbs and Showalter2015; Clarke and Pappu, Reference Clarke and Pappu2017). This involvement is likely due to their transient structures that can afford functional advantages over folded domains. For one, IDRs permit more rapid and efficient regulation compared to their folded counterparts (Uversky et al., Reference Uversky, Oldfield and Dunker2005; Sugase et al., Reference Sugase, Dyson and Wright2007). They can bind partners with broad selectivity and significant avidity; avidity refers to enhancing binding through many low-affinity poses in equilibrium (Wright and Dyson, Reference Wright and Dyson2015; Erlendsson and Teilum, Reference Erlendsson and Teilum2021). These low affinity interactions are frequently characterized by fast association and dissociation kinetics that enable switch-like changes in function (Wright and Dyson, Reference Wright and Dyson2015). For IDR-mediated PPIs, there is likely also a desolvation energetic advantage, as many IDPs have hydrophobic binding interfaces and undergo apolar desolvation upon binding to targets (Gibbs and Showalter, Reference Gibbs and Showalter2015).

IDRs are ubiquitous in myofilament proteins, which suggests their importance in muscle contraction and regulation (Best, Reference Best2017; Clarke and Pappu, Reference Clarke and Pappu2017). To quantify this propensity, we used PONDR (Obradovic et al., Reference Obradovic, Peng, Vucetic, Radivojac, Brown and Dunker2003) to predict disordered regions within the genes listed in section ‘Myofilament-associated protein with intrinsic disorder (MAPID)s’. These results are summarized in Fig. 3. We provide pie charts for each of the MAPIDs, categorized into thin/thick filaments(s), Z-disk, and miscellaneous, for which gray denotes folded regions versus red for IDRs. The MAPIDs we consider in this review are predicted to contain IDRs, though we recognize that some of these proteins are known to be folded as isolated (TNNC1) or co-assembled (ACTC1) proteins. These discrepant examples may just be due to inaccuracies of the PONDR algorithm. Alternatively, they may support the idea that all proteins are composed of both folded and disordered regions, but to varying degrees (Sormanni et al., Reference Sormanni, Piovesan, Heller, Bonomi, Kukic, Camilloni, Fuxreiter, Dosztanyi, Pappu, Babu, Longhi, Tompa, Dunker, Uversky, Tosatto and Vendruscolo2017). For instance, the predicted IDRs in TNNC1 all contain loops that resemble IDRs in the absence of Ca2+ (Fig. S1).

We next show how the IDR primary sequence affects the ensemble topology using an IDP state diagram developed by Pappu and coworkers (Das and Pappu, Reference Das and Pappu2013; Das et al., Reference Das, Ruff and Pappu2015; Holehouse et al., Reference Holehouse, Das, Ahad, Richardson and Pappu2017) (Fig. 3a). This diagram relies on two inputs, f + and f , that reflect the fractions of positively and negatively charged residues, respectively. These inputs classify IDRs into five ensemble states that differ in compactness (see the legend of Fig. 3 for detailed descriptions of these states). These state diagram coordinates allow us to estimate the approximate topology of the predicted IDRs (Figs 3b3e). For instance, for MYL7 we see that most of its IDRs fall within the R1 and R2 regions that are characterized as ‘premolten globules’ and the coexistence of globules and coils, respectively. However, one IDR is described as a swollen coil topology. Other MAPID examples feature IDRs that span diverse topologies. It is not yet understood if the distributions of topologies are by chance or are shaped by their respective roles in regulating myofilament function.

Post-translational modifications

PTMs provide a mechanism to rapidly, and often reversibly (Jideama et al., Reference Jideama, Crawford, Hussain and Raynor2006), perturb the ensemble from its native (unmodified) configuration. Just as the state diagrams show how an IDP's charge distributions determine its conformation ensemble topologies, they can be used to predict how PTMs may alter the native ensemble. Phosphorylation is a well-studied PTM from the broad variety of modifications found among MAPIDs. Phosphorylation is a form of chemical modification of tyrosines, serines and threonines that alters a given IDR's charge distributions, changes local/non-local electrostatic interactions (inter- and intrachain), and impacts solvation. Moreover, IDRs usually have multiple phosphorylation sites (Martin et al., Reference Martin, Holehouse, Grace, Hughes, Pappu and Mittag2016) that can amplify those changes. Human cardiac TnI, for instance, has 14 PTMs that have so far been characterized, of which the majority are phosphorylations (Biesiadecki and Westfall, Reference Biesiadecki and Westfall2019). These PTMs are believed to be the cornerstones of TnI's regulatory role on myofilament function in response to kinase activities (Biesiadecki and Westfall, Reference Biesiadecki and Westfall2019). Of these, a host of phosphorylation sites on the disordered N-terminus of TnI influence its conformation and are implicated in weakening the Ca2+ sensitivity of TnC (Cheng et al., Reference Cheng, Lindert, Kekenes-Huskey, Rao, Solaro, Rosevear, Amaro, McCulloch, McCammon and Regnier2014; Lindert et al., Reference Lindert, Cheng, Kekenes-Huskey, Regnier and McCammon2015; Zamora et al., Reference Zamora, Papadaki, Messer, Marston and Gould2016).

To illustrate this potential across the broad MAPID family, we collected phosphorylation sites within predicted IDRs from the PhosphoSitePlus database (Hornbeck et al., Reference Hornbeck, Zhang, Murray, Kornhauser, Latham and Skrzypek2015) (see section S1 for methodology). The statistics reported in Figs 3(b)–3(e) demonstrate that MAPID IDRs are rife with PTMs, for which only 5 out the 31 proteins (MYL3, MYL7, TNCC1, CRIP3, FHL2) do not yet report PTM sites in the predicted IDR regions. Genes with the most abundant PTMs include ANK2 and TTN with 119 and 338, respectively. We next show how these phosphorylation sites may affect the IDRs' ensemble properties using the IDR state diagrams (Figs 3b–3e). To mimic the negative charge brought by phosphorylation, we mutated the residues identified as PTM sites to glutamic acids, which altered the IDR charge distributions. The pink dots in Figs 3b–3e highlight the corresponding changes in ensemble topology upon phosphorylation. Since the phosphate groups modulate the intrinsic charge and charge density of the IDRs, the ensemble properties for the modified proteins are shifted relative to the unmodified states. For many examples, the modified versus unmodified proteins' distributions fall in different phases of the diagrams. As an example, the unmodified myosin heavy chains (MYH6 and MYH7) are distributed along the R2/R3 borders, which represent the coexistence of random coils and globules. The phosphorylation-modified states, however, are redistributed toward the R3/R4 border, suggesting an ensemble change to relatively extended swollen coils. Although it remains unclear which conformational ensembles are important for myosin heavy chain function, we speculate that topological changes may enable IDRs' influence on myofilament function to be tuned, such as to increase the likelihood of a PPI or to enhance its avidity.

Single nucleotide polymorphisms

The apparent importance of IDRs in myofilament function and PTM suggests that missense mutations in these regions similarly perturb myofilament function. To assess this potential, we retrieved pathogenic and likely pathogenic variants for each protein from the ClinVar database. We focused on the protein-changing variants, e.g. those resulting in missense mutations or deletions. Of the 36 proteins considered for this study, only 8 proteins did not have reported protein-changing variants. These include ABLIM1, CRIP3, FHL2, SYNPO2, TGM2, TMOD1, MYL7, and ENH2. We report in Fig. 3 disease-related mutations that reside in predicted IDRs. Similar to phosphorylation, missense mutations that alter an IDR's intrinsic charge have the potential to alter the IDR topology. Those that conserve charge could still impact native structure, depending on the side chain properties or backbone torsion angles (Jumper et al., Reference Jumper, Faruk, Freed and Sosnick2018). Here we find that potentially pathological missense mutations are found in most of the MAPID IDRs, which is consistent with observations made for proteins in general (Vacic and Iakoucheva, Reference Vacic and Iakoucheva2012). In contrast to the seemingly uniform PTM abundance across the core and associated protein sets, the disease-related mutations are more prevalent in the core proteins. This may imply that either disturbance of core protein functions presents a more obvious phenotype than the associated proteins, or simply that the core proteins have been genotyped to a greater degree.

General challenges in characterizing native IDR structures

Bioinformatic approaches generally determine whether an amino acid sequence would give rise to a disordered polypeptide, but do not in general predict their physicochemical properties. Many of these approaches rely on the observation that IDPs lack the hydrophobic core typical of globular proteins, and are instead enriched with charged and polar residues (Clarke and Pappu, Reference Clarke and Pappu2017). These residues in turn determine the local/non-local intrachain interactions, solvent/protein interactions, and ultimately IDP ensemble properties (Fig. 5). A key challenge for understanding these interactions and properties in IDRs is that their complete conformation ensembles must be resolved instead of a single conformational state. This is because IDRs have rugged potential energy surfaces that present numerous conformations in equilibrium that rapidly interchange. Relatedly, unlike globular proteins whose folding landscape is usually single-funneled, IDPs' energy landscapes are multi-funneled, which significantly challenges structure prediction tools such as AlphaFold (Strodel, Reference Strodel2021). Altogether, the large number of poorly resolved, functional conformations, and their equally important transition kinetics, pose fundamental challenges for the understanding of structure and function relationships among MAPIDs.

Fig. 5. (a) Major secondary structural elements present in folded proteins like actin (PDB 6KN8 (Yamada et al., Reference Yamada, Namba and Fujii2020)). (b) NMR structures of a MAPID, the MyBPC3 construct consisting of the M- and C2 domains (Michie et al., Reference Michie, Kwan, Tung, Guss and Trewhella2016), are shown as an example to illustrate the IDR ensemble. (c) Scaling of radius of gyration (R g) versus chain length for proteins at different states (Lazar et al., Reference Lazar, Martnez-Prez, Quaglia, Hatos, Chemes, Iserte, Mndez, Garrone, Saldao, Marchetti, Rueda, Bernad, Blackledge, Cordeiro, Fagerberg, Forman-Kay, Fornasari, Gibson, Gomes, Gradinaru, Head-Gordon, Jensen, Lemke, Longhi, Marino-Buslje, Minervini, Mittag, Monzon, Pappu, Parisi, Ricard-Blum, Ruff, Salladini, Skep, Svergun, Vallet, Varadi, Tompa, Tosatto and Piovesan2021). (d) The IDP phase-diagram for classifying IDPs into five structural states (R1–R5) based on the charge patterns (Holehouse et al., Reference Holehouse, Das, Ahad, Richardson and Pappu2017).

The folded and disordered continuum

Dividing proteins into ordered and disordered states is an oversimplification. Recent experiments, especially solution NMR techniques that can capture both conformations and their dynamics, suggest that protein ensembles are heterogeneous and contain both ordered and disordered components (Sormanni et al., Reference Sormanni, Piovesan, Heller, Bonomi, Kukic, Camilloni, Fuxreiter, Dosztanyi, Pappu, Babu, Longhi, Tompa, Dunker, Uversky, Tosatto and Vendruscolo2017). Similarly, IDRs can contain fragments with folded character. These are called residual structures and are of great importance to protein functions, including binding to protein partners (Wicky et al., Reference Wicky, Shammas and Clarke2017). It is increasingly recognized that proteins function as ensembles and even the so-called native states of folded proteins consist of multiple conformations (Gibbs and Showalter, Reference Gibbs and Showalter2015). Therefore, an IDR represents an extreme case of this paradigm (Gibbs and Showalter, Reference Gibbs and Showalter2015), which we illustrate with two examples. As one example, the switch peptide region of TnI (residues R148–K164) is disordered in the absence of a target, but adopts an alpha-helix upon binding Ca2+-activated TnC (Lindert et al., Reference Lindert, Cheng, Kekenes-Huskey, Regnier and McCammon2015; Cool and Lindert, Reference Cool and Lindert2021). Another example is the M domain of MyBPC3 (residues T255–R357), which retains a small tri-helix bundle that is believed to bind myosin (Howarth et al., Reference Howarth, Ramisetti, Nolan, Sadayappan and Rosevear2012; Michie et al., Reference Michie, Kwan, Tung, Guss and Trewhella2016; Singh et al., Reference Singh, McNamara and Sadayappan2021). Along these lines, loops or linkers between folded elements like helices (such as TnC's loops), and N/C termini (such as TnI) are frequently disordered (as has been predicted by PONDR in Fig. 3). The disorder inherent in these linkers can influence the function of folded domains (Sun and Kekenes-Huskey, Reference Sun and Kekenes-Huskey2022), as opposed to just serving a passive role in linking the folded domains.

Modifiers of IDR ensembles

Physicochemical properties of IDRs are encoded both by the number of covalently bonded amino acids and non-covalent properties bestowed by their side chains. Non-covalent interactions of an IDR with its environment help determine its properties. We discuss these factors in detail below.

Covalent interactions The ‘size’ of an IDR is often characterized by its radius of gyration (R g). Intuitively and generally, the longer the IDR's sequence, i.e. the number of covalently linked amino acids, the larger its R g. Analytic approximations for R g as a function of the numbers of amino acids are explained in section ‘Implicit/semi-analytic representations’. Independent of the non-covalent effects discussed below, covalent character can influence IDR properties in two ways. (1) Crosslinking of an IDR through disulfide bonds will reduce an ensemble's R g. (2) Differences in amino acid backbone distributions, often characterized by their ϕ/ψ (Ramachandran) angles, can shape the IDR conformation ensemble. Prolines and glycines, as an example, constrain and relieve constraints respectively (Huang and Nau, Reference Huang and Nau2003), on peptide backbone conformations. More generally, the ϕ and ψ dihedral angles assumed by contiguous amino acids support secondary structure inclusive of beta sheets, helices (α-helix/π-helix/3–10 helix), and turns, though non-covalent interactions arising from hydrogen bonding ultimately stabilize these structures.

Non-covalent interactions The abundance of polar amino acids that favorably interact with a polar solvent, water, relative to hydrophobic residues is a strong determinant of IDR character. These properties constitute non-covalent effects that simply dictate whether an amino acid prefers (thermodynamically speaking) to interact with the solvent or with the solute protein.

Hydrophobicity and hydrophilicity Well-folded proteins comprise hydrophobic residues that are thermodynamically disfavored to interact with polar solvents. For this reason, there is a thermodynamic driving force for hydrophobic residues to coalesce into compact domains that minimize polar solvent interactions. This occurs despite the loss of entropy that occurs during protein folding (Cheung et al., Reference Cheung, García and Onuchic2002). Conversely, polar residues favor interactions with both other polar residues and solvent. In many cases, since the enthalpy of a polar amino acid's interaction with water or other polar residues are similar, maximizing entropy favors unfolded states of the protein (Toal et al., Reference Toal, Verbaro and Schweitzer-Stenner2014). For this reason, IDRs tend to have an abundance of polar and charged amino acids (Lieutaud et al., Reference Lieutaud, Ferron, Uversky, Kurgan, Uversky and Longhi2016).

Electrostatic effects A solute's free energy of solvation depends on its charge, the solvent dielectric, and the presence of other charged solutes (Eq. (9)). Hence, the driving force for folding versus maintaining an intrinsically disordered, unfolded configuration will depend on the solute's environment and its physicochemical properties (Fig. 6). Factors of the environment include changes in ionic strength and temperature (see κ and T in Eq. (9)), as well as the presence of crowders or organelles that impact charges (q i) and dielectric constants ($\epsilon _{\rm r}$). Physicochemical properties can include reversible (protonation, phosphorylation, and N$\epsilon$-acetylation) and frequently irreversible (oxidation, glycation, and Nα-acetylation) chemical changes to a protein's structure (Vu et al., Reference Vu, Gevaert and De Smet2018; Narita et al., Reference Narita, Weinert and Choudhary2019).

Fig. 6. Schematic of transient and covalent influences on IDR structure. Ions can screen intramolecular electrostatics or directly coordinate with charged residues. pH affects the protonation state of ionizable residues (e.g. histidine). Phosphorylation introduces negative charges into the sequence. For crowding, many factors such as size and surface charges of crowders, and volume fraction affect IDR structure (Eq. (7)).

Ions Ionic strength changes follow the addition or removal of ionic species (monatomic species such as K+ and Cl, as well as molecules like ATP4−), denaturants like guanidinium chloride, or species that impact pH (discussed below). Ions primarily modulate IDR structural properties in two ways: they directly interact with IDR residues to modify net charges or screen intrachain electrostatic interactions (Uversky, Reference Uversky2009). In the former case, binding ions such as Zn2+ change the net charge of a residue or even ionically link residues to exert a direct impact on an IDR's secondary structure (Uversky, Reference Uversky2009). The dependence of these factors on ion types has been reported in Wicky et al. (Reference Wicky, Shammas and Clarke2017). Namely, the authors found that the residual structures of an IDR in its isolated state and the transition state for binding are sensitive to different types of ions, resulting in altered association and dissociation kinetics. Similarly, ions can bind to and interfere with the dynamics of mobile loops in proteins, as was shown for the binding of Ca2+ to titin's Ig domains (Kelly et al., Reference Kelly, Pace, Gage and Pfuhl2021). In addition, ions can screen electrostatic interactions, which is common in physiological conditions. The extent of this effect as a function of ionic strength can be estimated via the Debye–Huckel model (Eq. (9)). At high ionic strength, ions can have a ‘salting-out’ effect that reduces amino acid solubility. This effect is dependent on the type of ion (Hofmeister series) (Wohl et al., Reference Wohl, Jakubowski and Zheng2021).

pH The pH quantifies the balance of H+ and OH ions in solution. IDR ensemble properties can be altered by pH through changing the protonation of ionizable amino acids or ionic strength (Sun et al., Reference Sun, Blood, Atalay, Colli, Rankin, Barbara and Kekenes-Huskey2020). Protonation changes in turn alter an IDR's charge distribution and salt-bridges. Ionization state changes are most apparent in very acidic or basic conditions, where dramatic structural changes may be evident (Uversky, Reference Uversky2009). The PEVK repeats of titin exemplify this effect, as it shows a pH-dependent compaction of its structure (Manukian et al., Reference Manukian, Punch and Gage2022). An IDR prediction tool named DispHred has been developed to characterize disorder propensity as a function of pH (Santos et al., Reference Santos, Iglesias, Pintado, Santos-Surez and Ventura2020).

Crowding Crowders refer to molecular and subcellular material that vary in terms of charge density, volume, capacitance (membranes), and internal dielectric constants (Fig. 6). Crowders can modify an IDR ensemble (Banks et al., Reference Banks, Qin, Weiss, Stanley and Zhou2018), by altering the free energies of the unfolded, solvated states. In a dilute solution, configurational entropy is maximized (reducing the free energy) as a polymer disperses into the medium. In the presence of crowders, the space within which the polymer can disperse is reduced, which in turn reduces the entropy gain. Hence, crowding can shift the balance between folded and unfolded states, simply by reducing the free volume that a polymer can occupy. Along these lines, it is intriguing that the spacing between myofilaments changes occurs during contraction, which in principle could modulate the free volume available to MAPIDs (Irving et al., Reference Irving, Konhilas, Perry, Fischetti and de Tombe2000). An excellent review on the topic of how crowders influence protein thermodynamics was published by Zhou (Reference Zhou2008).

Post-translational modification (PTM) Post-translational modifications represent chemical (covalent) changes to protein structure that typically alter the non-covalent properties of an IDR. These can include increasing the size of an amino acid, its charge, and even modify inter-/intraprotein contacts. PTMs provide a reversible mechanism for rapid modulation of a protein's structure, protection against reactive species, and tagging for degradation or subcellular transport (Vu et al., Reference Vu, Gevaert and De Smet2018). IDRs usually have multiple phosphorylation sites (Martin et al., Reference Martin, Holehouse, Grace, Hughes, Pappu and Mittag2016), which provides a reversible means to alter a protein's charge distribution, local and non-local electrostatic interactions, and solvation thermodynamics (Martin et al., Reference Martin, Holehouse, Grace, Hughes, Pappu and Mittag2016). However, at least in some cases, global properties of IDRs such as the radius of gyration are insensitive to phosphorylation owing to compensatory changes in intrachain interactions (Martin et al., Reference Martin, Holehouse, Grace, Hughes, Pappu and Mittag2016). This suggests cell signaling may recognize local changes in sequence and charge, rather than changes in the global structure of the conformational ensemble. For example, many IDRs have SLIM regions that are used for binding protein targets. Since PTMs are abundant among IDRs and approximately 2000 SLIMs have already been identified (Lindorff-Larsen and Kragelund, Reference Lindorff-Larsen and Kragelund2021), we speculate that PTMs could be an important mechanism for toggling the availability of SLIMs for target binding.

A well-studied example of PTMs among MAPIDs include phosphorylation of TnI at Ser22/23 in its N-terminal IDR, which alters TnI binding to TnC (Hwang et al., Reference Hwang, Cai, Pineda-Sanabria, Corson and Sykes2014). Another includes MyBPC3, for which phosphorylation impacts myofilament contraction by altering its interactions with actin (Previs et al., Reference Previs, Mun, Michalek, Previs, Gulick, Robbins, Warshaw and Craig2016). In recent years, broad data sets of PTMs in MAPIDs have been identified through bottom-up (proteolytic digestion before mass spectrometry) (Kooij et al., Reference Kooij, Venkatraman, Kirk, Ubaida-Mohien, Graham, Faber and Van Eyk2014) and top–down (digestion-free) (Zabrouskov et al., Reference Zabrouskov, Ge, Schwartz and Walker2008) mass spectrometry approaches. Extensive reviews of PTM types are in the literature (Deribe et al., Reference Deribe, Pawson and Dikic2010; Prabakaran et al., Reference Prabakaran, Lippens, Steen and Gunawardena2012), but their impacts beyond phosphorylation and oxidation have not been extensively investigated in myofilament proteins, much less MAPIDs. This knowledge is critical for understanding both regulatory and dysregulated functions of proteins with IDRs (Uversky, Reference Uversky2014), including MAPIDs.

Challenges specific to MAPIDs

Many challenges have stymied the characterization of IDRs in MAPIDs. The staggering size of some myofilament proteins is one such challenge. Titin, as an example, consists of approximately 34 000 residues, of which nearly 35% of its primary sequence is predicted to be intrinsically disordered by PONDR (Obradovic et al., Reference Obradovic, Peng, Vucetic, Radivojac, Brown and Dunker2003) (Fig. 3). However, extensive simulations and experimental techniques that are commonly used to probe IDRs (Robustelli et al., Reference Robustelli, Piana, Shaw and Shaw2020; Chang et al., Reference Chang, Wilson, Karunatilleke, Moselhy, Karttunen and Choy2021; Ding et al., Reference Ding, Wang and Zhang2021) are best suited for much smaller proteins. Therefore, many studies of intact MAPIDs are limited by the large number of disordered residues, as well as the likely presence of multiple disordered regions within a single protein.

It is also recognized that the in vitro environment complicates the identification of IDRs. The myofilament is a complex structure consisting of dozens of proteins that function interdependently. As an example, actin has a well-folded structure in complex with troponin, and tropomyosin as shown in the crystal structure resolved by Yamada et al. (PDB 6KN8 (Yamada et al., Reference Yamada, Namba and Fujii2020)). At the same time, actin exhibits promiscuous target binding that is a signature of IDPs (Povarova et al., Reference Povarova, Uversky, Kuznetsova and Turoverov2014) and is also predicted to contain several IDRs (Turoverov et al., Reference Turoverov, Kuznetsova and Uversky2010; Povarova et al., Reference Povarova, Uversky, Kuznetsova and Turoverov2014). In addition, actin uses chaperones to fold onto the thin filament (Grantham, Reference Grantham2020). These observations suggest that the presence of other accessory proteins in an MAPID's environment can dictate the extent of its folding.

IDRs also often undergo disordered-to-ordered transitions, such as during a coupled binding and folding process (Sugase et al., Reference Sugase, Dyson and Wright2007). The C-terminal region of TnI provides an excellent example of this phenomenon. When the N-terminal domain of TnC is free of Ca2+ during diastole, its hydrophobic domain is closed and thus unavailable to bind the TnI C-terminal domain. This TnI domain therefore remains unbound and assumes an intrinsically disordered ensemble as determined via NMR (Julien et al., Reference Julien, Mercier, Allen, Fisette, Ramos, Lagüe, Blumenschein and Sykes2011). At saturating calcium during systole, TnC's hydrophobic domain is unveiled and supports binding of the TnI ‘switch’ peptide, during which the region folds into an alpha-helix (see Fig. 5a, e.g.). Modeling these processes is non-trivial and often suffers from force field inaccuracies and sampling insufficiency (Pietrek et al., Reference Pietrek, Stelzl and Hummer2020).

Lastly, there remains a challenge of relating computation-predicted IDR ensemble properties to experimental probes of the myofilament. A main reason for this is that there are spatial and temporal gaps in the resolution of experimental versus computational data sets. We refer to our experience with TnI binding to TnC, where ultimately the properties of TnI's conformational ensemble determine its experimentally observed affinity to TnC (Siddiqui et al., Reference Siddiqui, Tikunova, Walton, Liu, Meyer, de Tombe, Neilson, Kekenes-Huskey, Salhi, Janssen, Biesiadecki and Davis2016). Although multi-state models that take molecular-level IDR descriptions as inputs are helpful in predicting myofilament observables like binding affinity (Siddiqui et al., Reference Siddiqui, Tikunova, Walton, Liu, Meyer, de Tombe, Neilson, Kekenes-Huskey, Salhi, Janssen, Biesiadecki and Davis2016; Sun and Kekenes-Huskey, Reference Sun and Kekenes-Huskey2020), these models are often system-specific and difficult to generalize to other IDR-driven processes.

Part 2: computational modeling and characterization of MAPIDs

In this part, we overview the state-of-the-art computational methods used to characterize IDRs and IDPs. We organized this section as a potential workflow for an investigator to characterize a newly identified IDR. Hence, we divide these approaches into those targeting ensembles of IDRs in steady state (section ‘Computational methods for predicting conformation ensembles of isolated MAPIDs’), the kinetics of intramolecular dynamics (section ‘Computational methods for predicting intramolecular dynamics of MAPIDs’), and complex assembly (section ‘Computational methods for predicting the MAPID co-assembly’). In each section, we motivate the specific challenge for each type of approach with a brief synopsis of relevant experimental techniques, since excellent reviews of experimental methods for IDRs are readily available (see Gibbs and Showalter, Reference Gibbs and Showalter2015; Schramm et al., Reference Schramm, Bignon, Brocca, Grandori, Santambrogio and Longhi2019). These synopses are followed by detailed summaries of commonly used computational approaches and their applications to MAPIDs, when available.

Computational methods for predicting conformation ensembles of isolated MAPIDs

Problem and application

A primary goal in the characterization of proteins with IDRs is to describe their structure and physicochemical properties under equilibrium and steady-state conditions. We distinguish this challenge from that of assessing intraprotein kinetics and intermolecular assembly that we discuss in subsequent sections. Characterizing the steady-state properties of isolated IDRs typically entails describing the spatial extent of the conformation ensemble, the predominant conformations forming the ensemble, as well as the secondary, tertiary, and quaternary structure of the conformations. Also of interest is how the molecular and environmental factors summarized in section ‘Modifiers of IDR ensembles’ alter the ensembles' properties. These insights provide an initial foundation for understanding the biological role of MAPIDs. Such is the case for studies seeking to characterize the disordered switch region in TnI that primes its interaction with TnC (Lindert et al., Reference Lindert, Cheng, Kekenes-Huskey, Regnier and McCammon2015; Siddiqui et al., Reference Siddiqui, Tikunova, Walton, Liu, Meyer, de Tombe, Neilson, Kekenes-Huskey, Salhi, Janssen, Biesiadecki and Davis2016; Cool and Lindert, Reference Cool and Lindert2021). Equilibrium properties like the shape of an IDR's ensemble can impact physiologically important parameters such as myofilament length (nebulin (Wu et al., Reference Wu, Forbes and Wang2016)) and passive tension (titin PEVK motifs (Yamasaki et al., Reference Yamasaki, Berri, Wu, Trombitas, McNabb, Kellermayer, Witt, Labeit, Labeit, Greaser and Granzier2001)). The ensemble's shape also impacts the ability to assemble macromolecular structures or bind targets (discussed further in section ‘Computational methods for predicting the MAPID co-assembly’) to engage in PPIs, such as TnC binding to TnI (Metskas and Rhoades, Reference Metskas and Rhoades2015; Siddiqui et al., Reference Siddiqui, Tikunova, Walton, Liu, Meyer, de Tombe, Neilson, Kekenes-Huskey, Salhi, Janssen, Biesiadecki and Davis2016; Cool and Lindert, Reference Cool and Lindert2021).

Experimental techniques

The state-of-the-art for the experimental characterization of IDPs has been robustly reviewed (see, e.g. Gibbs and Showalter, Reference Gibbs and Showalter2015; Schramm et al., Reference Schramm, Bignon, Brocca, Grandori, Santambrogio and Longhi2019), therefore here we briefly introduce methods that are commonly used in tandem with computational approaches. Among these include SAXS, NMR, CD, hydrogen/deuterium exchange mass spectrometry (HDXMS), and FRET. Generally, SAXS provides information on gross protein shape and compactness, CD assesses the relative abundance of secondary structure content, NMR and HDXMS can monitor conformational dynamics, and FRET can probe quaternary structure or oligomerization (Schramm et al., Reference Schramm, Bignon, Brocca, Grandori, Santambrogio and Longhi2019; Metskas and Rhoades, Reference Metskas and Rhoades2020). In addition, cryo-EM and atomic force techniques have also been used to study IDP oligomerization and mechanical parameters (Karsai et al., Reference Karsai, Kellermayer and Harris2011; Mostofian et al., Reference Mostofian, McFarland, Estelle, Howe, Barbar, Reichow and Zuckerman2022). These methods can also be used in combination. Because the aforementioned methods are priorly used for probing equilibrium properties of IDRs, we extensively discuss the fundamentals of each technique in this section. Specific applications of these methods to the kinetics of IDR ensembles and kinetics are discussed briefly in subsequent sections of the review.

Size-exclusion chromatography

Size-exclusion chromatography is used to separate proteins by their sizes according to their elution time in a porous column. The elution time of a protein is related to its Stokes radius:

(1)$$R_{\rm s} = \displaystyle{{k_{\rm B}T} \over {6\pi \eta D}}$$

where η is the solvent's viscosity, D is the diffusion coefficient, k B is Boltzmann constant and T is temperature. Proteins with the same mass but larger Stoke radius will diffuse through the column instead of being captured by its porous interior and thus elute earlier than would be predicted from the protein mass alone (Schramm et al., Reference Schramm, Bignon, Brocca, Grandori, Santambrogio and Longhi2019). A larger Stokes radius is indicative of an unfolded protein relative to a folded protein of a similar size. This method has similar advantages and disadvantages relative to electrophoretic probes of mobility discribed next.

R g is a commonly used metric to assess the size of an IDP. Although the R g and the Stokes radius (R s) are typically measured by different methods (static scattering measurements for R g and dynamic light scattering for Stoke radius, respectively), they are intrinsically related (Tande et al., Reference Tande, Wagner, Mackay, Hawker and Jeong2001). For a hard sphere, R g/R s ≈ 0.77, and this ratio changes with the aspect ratio of the polymer. A rod-like polymer can have a ratio of up to 1.27 (Tande et al., Reference Tande, Wagner, Mackay, Hawker and Jeong2001). Size exclusion chromatography has been used quite extensively for MAPIDs. These applications include probes of titin's PEVK motifs (Duan et al., Reference Duan, DeKeyser, Damodaran and Greaser2006) and the disordered C-terminus of ankyrin B (Abdi et al., Reference Abdi, Mohler, Davis and Bennett2006), as well as assays using fesselin, a homologue of synaptopodin 2 (Khaymina et al., Reference Khaymina, Kenney, Schroeter and Chalovich2007), and FATZ-1 (Sponga et al., Reference Sponga, Arolas, Schwarz, Jeffries, Chamorro, Kostan, Ghisleni, Drepper, Polyansky, Ribeiro, Pedron, Zawadzka-Kazimierczuk, Mlynek, Peterbauer, Doto, Schreiner, Hollerl, Mateos, Geist, Faulkner, Kozminski, Svergun, Warscheid, Zagrovic, Gautel, Konrat and Djinovi-Carugo2021).

Gel electrophoresis

Gel electrophoresis leverages the principle that proteins with different molecular weights and net charges have different mobilities in an applied electric field. There are two kinds of gel electrophoresis: (1) native gel electrophoresis, and (2) SDS-PAGE (sodium dodecyl sulfate–polyacrylamide gel electrophoresis). For SDS-PAGE, the protein's intrinsic charges are masked by the sodium dodecyl sulfate. This masking enables all protein components to adopt similar charge-to-mass ratios, thus proteins are separated solely based on their molecular masses. In the native gel, proteins migrate in their native states; therefore, the migration is affected by the protein's intrinsic charge, size, and folding state (Arndt et al., Reference Arndt, Koristka, Bartsch and Bachmann2012).

Both strategies are useful tools for IDP studies. IDPs usually have larger apparent molecular masses than would be expected for a folded protein due to (1) their unique residue compositions, e.g. high percentages of charged residues and less hydrophobic residues, which make them bind less strongly to SDS, thus they migrate slower, (2) their unfolded states yield large Stokes radii, and (3) high proline content that leads to stiff conformations (Schramm et al., Reference Schramm, Bignon, Brocca, Grandori, Santambrogio and Longhi2019). Therefore, SDS-PAGE measurements of mobility are commonly used to probe whether a given protein has intrinsically disordered content.

The native gel is useful to infer the binding and folding of IDRs. As an example, Neirynck et al. probed how the CCT chaperone folds actin using native gel electrophoresis (Neirynck et al., Reference Neirynck, Waterschoot, Vandekerckhove, Ampe and Rommelaere2006). In that experiment, an alanine scan was performed on actin, whereby amino acids speculated to bind the chaperonin were mutated to alanine to assess if the binding interactions were compromised. The mutagenesis of amino acids that were directly involved in chaperone binding was identified by their impact on the electrophoresis of the variant actin relative to wild-type (Neirynck et al., Reference Neirynck, Waterschoot, Vandekerckhove, Ampe and Rommelaere2006). A primary advantage of gel electrophoresis for assessing IDR characteristics is that it is a straightforward and inexpensive technique to perform. The main disadvantages are that these analyses do not directly provide structural insights and the migration-based assessments of IDRs are qualitative. There are other factors such as detergent binding that can also affect migration (Rath et al., Reference Rath, Glibowicka, Nadeau, Chen and Deber2009).

Circular dichroism (CD)

CD is a technique that relies on the polarization of incident light through a sample. In the context of protein structure determination, secondary structures such as α-helices and β-sheets (see Fig. 5) rotate light at different angles. Hence, the secondary structure content can be monitored in IDPs to determine unfolded to folded transitions or partially folded domains in an IDR. The primary advantage is that CD can work with small concentrations of proteins with wide-ranging sizes. A disadvantage is that CD spectra of β-sheet structures are quite broad and overlap with α-helices (Micsonai et al., Reference Micsonai, Wien, Kernya, Lee, Goto, Réfrégiers and Kardos2015), thus limiting the analysis to qualitative assignment of secondary structures. CD has been used for probes of many MAPIDs, including titin's PEVK motifs (Ma and Wang, Reference Ma and Wang2003; Duan et al., Reference Duan, DeKeyser, Damodaran and Greaser2006), ankyrin B's C-terminus (Abdi et al., Reference Abdi, Mohler, Davis and Bennett2006), tropomodulin's intrinsically disordered N-terminus (Greenfield et al., Reference Greenfield, Kostyukova and Hitchcock-DeGregori2005), and for FATZ-1 (Sponga et al., Reference Sponga, Arolas, Schwarz, Jeffries, Chamorro, Kostan, Ghisleni, Drepper, Polyansky, Ribeiro, Pedron, Zawadzka-Kazimierczuk, Mlynek, Peterbauer, Doto, Schreiner, Hollerl, Mateos, Geist, Faulkner, Kozminski, Svergun, Warscheid, Zagrovic, Gautel, Konrat and Djinovi-Carugo2021). Other examples are listed in Table 1.

Table 1. Brief summary of reported experimental and computational studies on MAPIDs

Experimental methods and their observables supporting intrinsic disorder presence are briefly explained. EM: missing density. X-ray: unresolved structure or high B-factor. CD: little to no secondary structure. NMR: secondary structure and IDP-related observables (Brutscher et al., Reference Brutscher, Felli, Gil-Caballero, Hošek, Kümmerle, Piai, Pierattelli, Sólyom, Felli and Pierattelli2015). FRET: ensemble dimensions/dynamics, population heterogeneity. SAXS: compaction (R g) (Gibbs and Showalter, Reference Gibbs and Showalter2015). AFM: mechanical properties. Gel analysis: larger Stokes radius, slower migration (Schramm et al., Reference Schramm, Bignon, Brocca, Grandori, Santambrogio and Longhi2019).

Small-angle X-ray scattering (SAXS)

SAXS is an approach that measures the scattering of X-rays upon colliding with a solute in solution. In most applications, SAXS data are reported in Guinier plots that summarize the intensity of scattered X-rays versus the scattering vector. The radius of gyration (R g) is readily obtained from this analysis, which lends itself to the assessment of IDR sizes (Zheng and Best, Reference Zheng and Best2018):

(2)$$\ln \displaystyle{{I( q) } \over {I( 0) }} = ( {-}R_{\rm g}^2 /3) q^2$$

where q is the scattering vector, I(q) is the SAXS intensity at the scattering vector, q. I(0) is the intensity when q = 0. The I(q) profile from the Guinier plot also exhibits profiles that can indicate a protein ensemble's shapes, such as spherical versus oblong (Grupi and Haas, Reference Grupi and Haas2011). A popular SAXS-based IDR ensemble generation strategy, called the ‘ensemble optimization method (EOM)’ (Bernadó et al., Reference Bernadó, Mylonas, Petoukhov, Blackledge and Svergun2007), has been developed for which conformations are picked from a pool of conformations randomly generated by simulation to match SAXS data. Advantages of this method are that it is a label-free technique that can yield structural information for arbitrarily large proteins and macromolecular complexes. A chief limitation of the method is that access to a synchrotron is needed for an X-ray source. Examples of MAPIDs characterized by SAXS include FATZ-1 (Sponga et al., Reference Sponga, Arolas, Schwarz, Jeffries, Chamorro, Kostan, Ghisleni, Drepper, Polyansky, Ribeiro, Pedron, Zawadzka-Kazimierczuk, Mlynek, Peterbauer, Doto, Schreiner, Hollerl, Mateos, Geist, Faulkner, Kozminski, Svergun, Warscheid, Zagrovic, Gautel, Konrat and Djinovi-Carugo2021) and the M-domain of MyBPC3 (Michie et al., Reference Michie, Kwan, Tung, Guss and Trewhella2016).

Fluorescence spectroscopy

The intrinsic fluorescence of aromatic residues, namely tryptophan and to a far lesser extent tyrosine and phenylalanine, can be measured using fluorescence spectroscopy. Changes in fluorescence indicate differences in the amino acids' environment (Ghisaidoobe and Chung, Reference Ghisaidoobe and Chung2014), such as due to folding or the binding of targets. For instance, tryptophans buried inside the hydrophobic core of folded proteins are brighter and blue-shifted relative to those exposed to aqueous solvent. Upon solvent exposure of tryptophans due to unfolding, the emission spectrum red-shifts and dims (Khaymina et al., Reference Khaymina, Kenney, Schroeter and Chalovich2007; Yang et al., Reference Yang, Xiao, Zhao and Wu2015). The advantage of this technique is that there is no need to introduce extrinsic fluorescent probes and generally fluorimeters are inexpensive. Disadvantages include the low extinction coefficient that results in dim signals and ambiguity when more than one fluorescent species is present. Intrinsic fluorescence has been applied to synaptopodin 2's homolog fesselin to show that the fluorescent signals of the tryptophan residues have maxima around 334 nm, suggesting that those residues are solvent-exposed and therefore the protein's native state is unfolded (Khaymina et al., Reference Khaymina, Kenney, Schroeter and Chalovich2007).

Förster resonance energy transfer (FRET). FRET is another fluorescence-based technique frequently used in protein structure characterization. The method relies on stimulating a donor fluorophore at its absorption wavelength, after which it emits at longer wavelengths. An acceptor fluorophore can absorb the energy and emit the energy at an even longer wavelength. The strength of this energy transfer decays as I(r) = I(0)/r 6, where r is the distance between donor and acceptor fluorophores. Given this rapid spatial decay of the intensity, FRET is commonly used to measure 1–10 nm distances between probes. An additional advantage is that the technique can use more than one pair of probes (Lee et al., Reference Lee, Moran-Gutierrez and Deniz2015). A disadvantage of the method is that a chemical or protein construct, such as green fluorescence protein, is typically introduced to the protein of interest, which can alter its structure. A study of IDRs in troponin I made extensive use of FRET to resolve the positioning of their conformations relative to folded troponin C structures (Metskas and Rhoades, Reference Metskas and Rhoades2015).

Nuclear magnetic resonance spectroscopy (NMR)

NMR-based techniques have been extensively used to determine the structures of modestly sized (commonly <30 kD (Xu et al., Reference Xu, Zheng, Fan and Yang2006)) isolated proteins and protein/protein complexes at Angstrom-level resolutions that can rival X-ray crystallography. The list of NMR-based techniques amenable to probing IDPs is ever-growing (reviewed in Schneider et al., Reference Schneider, Blackledge and Jensen2019; Dyson and Wright, Reference Dyson and Wright2021). In short, NMR leverages the interaction between an applied magnetic field and the nuclear spins of atoms. The applied magnetic field splits into spectral lines, or energy levels, of different energies (Zeeman effect). Those energies may shift depending on their local environment (chemical shift). These properties enable one to assign peaks in NMR spectra to specific nuclei types in functional groups (typically hydrogen, carbon, and nitrogen), as well as determine their proximity to other functional groups such as aromatic rings. 2D heteronuclear NMR broadens this approach by exciting different nuclei and accounts for the relaxation times for coupled spins, which improves the ability to resolve and assign peaks.

Excited nuclear spins can be transferred to nearby, coupled nuclear spins (nuclear overhauser effect, NOE), which enables the determination of distances between coupled nuclei. This approach is generally more readily applied to folded proteins. In complement, residual dipolar couplings (RDC)s, which measure the angles between partially ordered magnetic nuclei and the magnetic field (Prestegard et al., Reference Prestegard, Bougault and Kishore2004), are also increasingly used in IDP experiments (Fisher and Stultz, Reference Fisher and Stultz2011). In addition, order parameters that report on the mobility of bond vectors, such as N-H, can also be measured from spin relaxation rates. Overall, NMR spectroscopy offers a wide-array of techniques that can be used to probe both steady-state and dynamic properties of protein conformation ensembles in IDPs (Schneider et al., Reference Schneider, Blackledge and Jensen2019; Dyson and Wright, Reference Dyson and Wright2021). Disadvantages of these approaches include the need for large amounts of protein, sample preparation such as labeling that can vary considerably in cost, and expensive instrumentation.

NMR techniques for IDPs and proteins with IDRs have been applied to several MAPIDs. Ma et al. used 2D H-H NMR (TOCSY and ROESY) to probe the residual structure of PEVK motifs in titin and how the disordered content changes with temperature and ionic strength (Ma and Wang, Reference Ma and Wang2003). Other applications used 15N-1H HSQC to investigate the secondary structure content of tropomodulin's N-terminal IDR (Greenfield et al., Reference Greenfield, Kostyukova and Hitchcock-DeGregori2005) and its binding site for tropomyosin (Greenfield et al., Reference Greenfield, Kostyukova and Hitchcock-DeGregori2005; Kostyukova et al., Reference Kostyukova, Hitchcock-DeGregori and Greenfield2007). 2D 13C-13C solid-state NMR analyses of $C_\alpha /C_\beta$ peaks of serine and alanine residues in the desmin head domain revealed that approximately 80% of serine and 40% alanine residues are in β-strand conformations, respectively (Zhou et al., Reference Zhou, Lin, Kato, Mori, Liszczak, Sutherland, Sysoev, Murray, Tycko and McKnight2021b), suggesting the presence of both structured and disordered elements (Zhou et al., Reference Zhou, Fleming, Lange, Hessel, Bogomolovas, Stronczek, Grundei, Ghassemian, Biju, Borgeson, Bullard, Linke, Chen, Kovermann and Mayans2021a). Applications of NMR to MyBPC3 via 2D 1H-15N HSQC show that its M-domain consists of an N-terminal IDR and a C-terminal folded subdomain (Howarth et al., Reference Howarth, Ramisetti, Nolan, Sadayappan and Rosevear2012). Heteronuclear NOE values of each amide group and 15N T2 relaxation suggest that the linker in the M-domain of MyBPC3 is highly dynamic and disordered (Michie et al., Reference Michie, Kwan, Tung, Guss and Trewhella2016).

Other techniques

There is an extensive collection of less-commonly used techniques for probing IDR structure. HDXMS is one such example that measures the extent to which a proton from a deuterated solvent is exchanged with those of the protein. Amino acids that are exposed to the solvent will have higher deuterated protein content than those that are well-buried and isolated from the solvent. Furthermore, highly mobile regions will exchange more rapidly than rigid regions (Fanning et al., Reference Fanning, Jeselsohn, Dharmarajan, Mayne, Karimi, Buchwalter, Houtman, Toy, Fowler, Han, Lainé, Carlson, Martin, Nowak, Nwachukwu, Hosfield, Chandarlapaty, Tajkhorshid, Nettles, Griffin, Shen, Katzenellenbogen, Brown and Greene2018). The proton content is assessed via mass spectrometry and was recently used to characterize disordered linkers in titin (Zhou et al., Reference Zhou, Fleming, Lange, Hessel, Bogomolovas, Stronczek, Grundei, Ghassemian, Biju, Borgeson, Bullard, Linke, Chen, Kovermann and Mayans2021a).

AFM is commonly used to investigate the extensibility of proteins under a traction force. Unfolded proteins or IDRs exhibit more pliant force–length relationships relative to folded proteins. This technique has been applied to titin's PEVK motifs (Linke et al., Reference Linke, Kulke, Li, Fujita-Becker, Neagoe, Manstein, Gautel and Fernandez2002) to discriminate between folded and unfolded regions, as well as MyBPC3 (Previs et al., Reference Previs, Mun, Michalek, Previs, Gulick, Robbins, Warshaw and Craig2016).

X-ray crystallography is used to probe well-folded proteins with unique conformations. While IDRs are generally not amenable to this technique, those that undergo unfolded to folded transitions upon binding a target can be resolved. An example application resolved the autoinhibited state of ankyrin, in which its IDR tail was folded (Chen et al., Reference Chen, Li, Wang, Wei and Zhang2017). Similarly, X-ray crystallography was used to reveal the complex between α-actinin and FATZ-1 (myozenin-1) fragments that contain IDRs (Sponga et al., Reference Sponga, Arolas, Schwarz, Jeffries, Chamorro, Kostan, Ghisleni, Drepper, Polyansky, Ribeiro, Pedron, Zawadzka-Kazimierczuk, Mlynek, Peterbauer, Doto, Schreiner, Hollerl, Mateos, Geist, Faulkner, Kozminski, Svergun, Warscheid, Zagrovic, Gautel, Konrat and Djinovi-Carugo2021). Lastly, for transglutaminase 2, X-ray crystallography was used to infer that this protein has several IDRs, based on missing structural information in the loop regions (Pinkas et al., Reference Pinkas, Strop, Brunger and Khosla2007). Other less commonly used techniques include electron paramagnetic resonance, proteolytic degradation, and Fourier transform infrared spectroscopy (Uversky, Reference Uversky and Renaud2020).

Computational approaches

Computational approaches are important tools that complement experiments in revealing molecular bases for IDR ensembles. Key computational approaches are described below.

Bioinformatics approaches

An initial objective in IDR studies is to identify regions in protein sequence data that may be disordered. Significant advancements have been made in the area of IDP and IDR prediction from amino acid sequenc alone. IDPs feature unique sequence patterns including the enrichment of charged and polar residues relative to globular proteins. This impacts their ensemble compactness based on the charge distribution in the sequence (Das et al., Reference Das, Ruff and Pappu2015). They also tend to contain more serine, glycine, and proline residues (Lieutaud et al., Reference Lieutaud, Ferron, Uversky, Kurgan, Uversky and Longhi2016), of which the latter two impose different torsional constraints on the protein backbone relative to the other amino acids (Krieger et al., Reference Krieger, Moglich and Kiefhaber2005). Additionally, IDRs tend to be depleted of aromatic residues (F, W, and Y) (Povarova et al., Reference Povarova, Uversky, Kuznetsova and Turoverov2014) and some hydrophobic amino acids (I, L, and V) (Lieutaud et al., Reference Lieutaud, Ferron, Uversky, Kurgan, Uversky and Longhi2016). These sequence trends lend themselves to predicting the IDP propensity based on amino acid identities. Currently, there are tens of web servers or stand-alone tools for IDP prediction based on physicochemical properties, templates, metadata, and machine learning (Liu et al., Reference Liu, Wang and Liu2019). Others focus on the detection of SLIMs important to binding IDRs to partnering proteins (Lindorff-Larsen and Kragelund, Reference Lindorff-Larsen and Kragelund2021). These tools have comparable performance, but can be dataset dependent (Liu et al., Reference Liu, Wang and Liu2019).

There are also techniques for predicting structures for IDPs. An IDP's ensemble can be explicitly generated from sequence via tools like flexible-meccano (Ozenne et al., Reference Ozenne, Bauer, Salmon, Huang, Jensen, Segard, Bernad, Charavay and Blackledge2012), which uses basic conformational potentials derived from coil statistics together with user-customized potentials. The tool directly generates IDP conformations which, with proper user-customized potentials, can match NMR and SAXS measured IDP ensemble properties. Qualitative attributes of IDPs structural states, such as compactness and coil/globule properties, can be estimated from its sequence via the IDP diagram from the Pappu lab (Holehouse et al., Reference Holehouse, Das, Ahad, Richardson and Pappu2017). It is important to note that although being termed ‘intrinsically disordered’, IDPs commonly adopt partially folded structures. These residual structures can serve as molecular recognition fragments important to function (Gibbs and Showalter, Reference Gibbs and Showalter2015). However, predicting these residual structures is challenging as mainstream tools such as DSSP, STRIDE, and KAKSI show significant discrepancies (Zhang and Sagui, Reference Zhang and Sagui2015).

Given their reasonable accuracy, inexpensive computational expense, and user-friendly features, IDP prediction tools are routinely used to gain insight into structural disorder in myofilament proteins, especially for those whose structural information is scarce. As one example, bioinformatics approaches have been used to predict the IDP regions of different TnI isoforms (Hoffman et al., Reference Hoffman, Blumenschein and Sykes2006) and concluded that the N- and C-termini of TnI are intrinsically disordered across all isoforms. A similar bioinformatic method was applied to the intact Tn complex (TnI, TnC, and TnT) and disease-associated variants (Na et al., Reference Na, Kong, Straight, Pinto and Uversky2016). This study revealed that all three components have IDRs, even the well-folded TnC. Furthermore, the authors conclude that some disease-associated mutants reduce the disorder of the Tn complex, due to localized disorder-to-order transitions (Na et al., Reference Na, Kong, Straight, Pinto and Uversky2016). An additional example includes a study utilizing two predictors that reveal that the C-terminus of desmin is an IDR (Anbo et al., Reference Anbo, Sato, Okoshi and Fukuchi2019). This IDR harbors the binding site for alphaB-crystallin (CRYAB) according to IDEAL (Intrinsically Disordered proteins with Extensive Annotations and Literature) annotations (Fukuchi et al., Reference Fukuchi, Sakamoto, Nobe, Murakami, Amemiya, Hosoda, Koike, Hiroaki and Ota2012; Anbo et al., Reference Anbo, Sato, Okoshi and Fukuchi2019).

Implicit/semi-analytic representations

Polymer theory can provide insights into IDP ensemble properties, such as the end-to-end distance probability p(r). The most basic polymer model ignores interactions between monomers and assumes they are freely jointed (ideal chain model). This model yields quantities like the mean squared end-to-end distance, ⟨R 2⟩ and radius of gyration, R g, as a function of the number of monomers (amino acids), N:

(3)$$\langle R^2\rangle = Nb^2$$
(4)$$\langle R_{\rm g}^2 \rangle = \displaystyle{{\langle R^2\rangle } \over 6}$$

where b, the Kuhn length, reflects the stiffness and local interactions. Other models can avoid overlapping monomers (worm-like chain (WLC) and excluded volume chain models) (Milstein and Meiners, Reference Milstein, Meiners and Roberts2013). The WLC uses the persistence length, ζ, which is approximately b/2, to give the mean squared end-to-end distance:

(5)$$\langle R^2\rangle =2\zeta L_0 \left[1-{{\zeta}\over{L_0}}(1-\exp(-L_0/\zeta)) \right]$$

where L 0 = NL b with L b and N representing the length of an amino acid and the number of amino acids, respectively. These models, when coupled with FRET experiments, are particularly useful for revealing chain compactness and dynamics of IDRs (reviewed in Schuler et al., Reference Schuler, Soranno, Hofmann and Nettels2016).

To account for intramolecular steric effects and interactions between monomers and solvents, the excluded volume concept was introduced in Zimm et al. (Reference Zimm, Stockmayer and Fixman1953). This led to the Flory theory that quantifies competition between monomer/monomer and monomer/solvent interactions in determining compactness. This model accounts for excluded volume and entropic contributions in estimating the free energy of a polymer. It yields an approximation of the form:

(6)$$\langle R\rangle \approx N^{\nu}$$

where ν is a scaling parameter (ν = 0.6 is common for disordered chains) (Grupi and Haas, Reference Grupi and Haas2011). Treatments based on Flory–Huggins theory have been used to fit radii of gyration as a function of a solution's crowded volume fraction and the overlap of disordered chains (Soranno et al., Reference Soranno, Koenig, Borgia, Hofmann, Zosel, Nettels and Schuler2014). Revised polymer models that explicitly account for charge distributions and monomer interactions, such as short-range repulsion and long-range electrostatics interactions, are also available to better describe IDP ensemble properties (reviewed in Ghosh et al., Reference Ghosh, Huihui, Phillips and Haider2022).

Related to the concept of excluded volume is that of crowding, which refers to the mixing of a solute in a solution composed of solvent and other biomolecules. This is appropriate for the cell cytoplasm, which has a free volume fraction of roughly 0.78 (Kekenes-Huskey et al., Reference Kekenes-Huskey, Scott and Atalay2016). Along these lines, a theoretical framework has been developed to predict the scaling of an IDP's compactness (⟨Rg⟩) with the free volume fraction ϕ in the presence of hard sphere crowders (Kang et al., Reference Kang, Pincus, Hyeon and Thirumalai2015):

(7)$$\langle R_{\rm g}( \phi ) \rangle = \langle R_{\rm g}( 0) \rangle f( x) $$
(8)$$x\approx \langle R_{\rm g}( 0) \rangle /( 4\pi /3) ^{1/3}\sigma _{\rm c}\phi ^{{-}1/3}$$

where ⟨Rg(0)⟩ is an IDR's native compactness in crowder-free solutions and σ c is the radius of the crowder. While the exact solution of the scaling factor f(x) is not defined exactly, it is bounded by $x{\rm \mathbin{\lower.3ex\hbox{$\buildrel\prec\over {\smash{\scriptstyle\sim}\vphantom{_x}}$}}\ {\cal O}}( 1 )$ and $x\gg {{\rm \cal O}}( 1 )$ , which correspond to a decreasing ⟨Rg⟩ and a coil-to-globule transition as ϕ increases, respectively.

Electrostatic interactions can also dramatically alter the size of a polymer. Here it is useful to define the electrostatic potential, V ele, between two charged spheres representing residues i and j that are separated by distance r ij with point charges q i and q j via the Debye–Huckel model (Chu and Wang, Reference Chu and Wang2019):

(9)$$V_{ele} = K_{coul}B( \kappa ) \sum {\displaystyle{{q_iq_j\exp ( {-}r_{ij}\kappa ) } \over {{\rm \epsilon }_rr_{ij}}}} $$

In this expression, κ ≡ λ −1, where λ represents the Debye length, the distance over which electrostatic effects are most strongly screened by electrolytes. The Debye length is formally defined as $\lambda = \sqrt {\epsilon _rk_{\rm B}T/\sum _in_i^0 q_i^2 }$, for which the denominator reflects the solution's ionic strength. B(κ) is an approximate constant for a dilute electrolyte concentration. While it is intuitive that like- and unlike-charged residues will repel and attract one another, it is evident from Eq. (9) that the ionic strength will dictate the strength of these interactions. Namely, high ionic strength attenuates electrostatic interactions, which thereby influences IDR compactness.

The IDRs predicted in Fig. 3 either cap the N- or C-terminus of a well-folded protein or link two well-folded domains. Here, polymer theoretic models can be used to predict an ensemble's distribution about a folded protein or domain anchor. An example model from Van Valen et al. (Reference Van Valen, Haataja and Phillips2009) provides an analytic form for a binding domain linked to an anchor by an unstructured region. Here, the effective concentration, c eff, is estimated via a Gaussian chain model (Van Valen et al., Reference Van Valen, Haataja and Phillips2009):

(10)$$c_{{\rm eff}}( R ) = \left({\displaystyle{3 \over {4\pi \xi L}}} \right)^{3/2}\exp \left({-\displaystyle{{3R^2} \over {4\xi L}}} \right)$$

where R is a user-provided distance between the polymer's two ends and L is the linker length. The effective concentration of the untethered domain at a given R can be used to estimate its likelihood of binding a target (see section ‘Implicit and semi-analytic representations’ Eq. (56); Siddiqui et al., Reference Siddiqui, Tikunova, Walton, Liu, Meyer, de Tombe, Neilson, Kekenes-Huskey, Salhi, Janssen, Biesiadecki and Davis2016).

Explicit representations

All-atom simulations Molecular simulations are commonly used to sample IDP conformations constituting an ensemble. Simulation-generated conformations enable direct interpretation of an IDP's structure–function relationships. Molecular simulation methods can be categorized into Monte Carlo and MD approaches. These forms of simulations are all based on a Hamiltonian, ${\cal H}$, for estimating the energy associated with N particles, which is given by the sum of the system's potentials and kinetic energies. These particles can be linked through bonding potentials. For example, the bonded potential of the Amber force field (Weiner et al., Reference Weiner, Kollman, Case, Singh, Ghio, Alagona, Profeta and Weiner1984) consists of three energy terms representing bond, angle, and dihedral terms:

(11)$$\eqalign{V_{bonded} = & \sum\limits_{i\in bonds} {k_{b, i}} ( l_i-l_{i, 0}) ^2 \cr \quad + & \sum\limits_{i\in angles} {k_{a, i}} ( \theta _i-\theta _{i, 0}) ^2 \cr \quad + & \sum\limits_{i\in torsions} {\sum\limits_n {V_{n, i}} } [ 1 + \cos ( n\omega _i-\gamma _i) ] } $$

where l i,0, θ i,0, and γ i are the bond length, angle value, and dihedral phase angle at equilibrium, k b,i, k a,i, V n are the force constants, and n in the torsion term refers to the multiplicity. The potentials associated with non-bonded interactions including van der Waals (vdW) and electrostatics are commonly given by:

(12)$$\eqalign{V_{non-bonded} =& V_{ele} + V_{vdw} \cr =& \sum\limits_i {\sum\limits_{\,j > i} {\displaystyle{{q_iq_j} \over {{\rm \epsilon }r_{ij}}}} } \cr \quad +& \sum\limits_i {\sum\limits_{\,j > i} {\displaystyle{{A_{ij}} \over {r_{ij}^{12} }}} } -\displaystyle{{B_{ij}} \over {r_{ij}^6 }}} $$

where r ij is the separation between atoms (beads) i and j, q i, q j, are point charges, and A ij, B ij are vdW constants of atoms i and j.

The parameters for the expressions are defined in a force field. Force fields have generally been optimized for well-folded proteins, thus early applications to IDPs presented artifacts including overpredictions of compaction and secondary structure formation (Henriques et al., Reference Henriques, Cragnell and Skepo2015). As simulations of IDPs matured, corrections to these force fields resulted in versions such as the Amber ff99SB-ILDN (Lindorff-Larsen et al., Reference Lindorff-Larsen, Piana, Palmo, Maragakis, Klepeis, Dror and Shaw2010) that mitigate these effects to a certain extent.

Monte Carlo and lattice simulations Monte Carlo (MC) algorithms utilize random displacements of molecular coordinates to identify the most thermodynamically favorable conformations. MC approaches define a Hamiltonian similar to those in Eq. (11) comprising bonded and non-bonded terms. At each iteration of the algorithm, a new system is generated from the current state (or initial state) by randomly displacing its coordinates. For the Metropolis MC algorithm, if the new system $\{ {\vector x} \} {\rm ^{\prime}}$ yields a total energy that is lower than that of the previous state $\{ {\vector x} \}$, the new state is stored as the current state. Otherwise, a new state is accepted if (Earl and Deem, Reference Earl and Deem2008):

(13)$$N_{unif}{\rm ( 0, \;1) \lt\; }p$$
(14)$${\rm with}\quad p = \exp ( -\beta [ U( \{ {\vector x}\} {\rm ^{\prime}}) -U( \{ {\vector x}\} ) ] ) $$

where $U( {\{ {\vector x} \} {\rm^{\prime}}} )$ and $U( {\{ {\vector x} \} } )$ are the energies of the new and previous state, respectively, and β = 1/k B T is the thermal energy of the system. The random displacements can be constrained to a uniformly spaced grid (lattice simulation) or assume continuous distributions. Metropolis MC simulations were recently used to show that the conformation ensemble of an 81 amino acid IDP transcription factor Ash1 from Saccharomyces cerevisiae in its native and phosphorylated had similar radii of gyration (Martin et al., Reference Martin, Holehouse, Grace, Hughes, Pappu and Mittag2016).

The simulation process accumulates an ensemble of conformations, e.g. ${\vector X} = \{ {{\{ {\vector x} \} }_1, \;\;{\{ {\vector x} \} }_2, \;\ldots , \;\;{\{ {\vector x} \} }_n} \}$, which can be used to compute metrics such as an ensemble-averaged radius of gyration via:

(15)$$\langle R_{\rm g}\rangle = \displaystyle{1 \over {\vert {\vector X}\vert }}\sum\limits_{n\in {\vector X}} {\sqrt {\displaystyle{{\sum\limits_i \vert \vert {\vector r}_i\vert \vert ^2m_i} \over {\sum\limits_i {m_i} }}} } $$

where ${\vector r}_i$ and m i are the position and mass of atom i, respectively. Because there is an energy associated with each conformation from evaluating Eqs. (11) and (12) (above), one can compute ensemble-averaged, thermodynamic quantities such as the free energy, ⟨G⟩, using:

(16)$$\langle G\rangle = \displaystyle{1 \over {\vert {\vector X}\vert }}\sum\limits_{\vector i} G ( \{ {\vector x}\} _i) $$

where $G( {{\{ {\vector x} \} }_i} )$ is the free energy for each simulated conformation, such as from molecular mechanics calculations.

Molecular dynamics (MD) simulations MD entails minimization and displacement of atoms according to forces, based on Eqs. (11) and (12). Minimization techniques including steepest descent and conjugate gradients move particle positions according to the gradient of the potential energy surface (PES) until the system energy is minimized (often local):

(17)$$\{ {\vector x}\} _{{N}^{\prime}} = \{ {\vector x}\} _N-\gamma \nabla U( \{ {\vector x}\} _N) $$

where $\{ {\vector x} \} _N^{\rm ^{\prime}}$ and $\{ {\vector x} \} _N$ are the new and current configurations consisting of N particles, respectively. $\nabla U( {{\{ {\vector x} \} }_N} )$ is the energy gradient at the current system's configuration, and γ is the step size along the gradient. This is done in order to reconcile bond and non-bond energies that are incompatible with the force field. However, for IDRs, minimization is generally likely to yield a single, local minimum despite many states sharing similar energies.

Dynamic approaches temporally evolve a system by integrating Newton's equations of motion for each atom:

(18)$$F( \{ {\vector x}\} ) = {-}\nabla U( \{ {\vector x}\} ) $$
(19)$$\displaystyle{{dv} \over {dt}} = \displaystyle{{F( \{ {\vector x}\} ) } \over m}$$

where $F( {\{ {\vector x} \} } )$ is the force acting on the system, v and m are the atom's velocity and mass. During the simulation, v and $\{ {\vector x} \}$ are repeatedly updated at every time step δt:

(20)$$v_{t + \delta t} = v_t + \delta tF( \{ {\vector x}\} _t) /m$$
(21)$$\{ {\vector x}\} _{t + \delta } = \{ {\vector x}\} _t + \delta tv_t$$

An initial guess for the system's coordinates ($\{ {\vec{r}( 0 ) } \}$) may come from the Protein Data Bank. Particle velocities, v i, are randomly drawn from a Maxwell–Boltzmann velocity distribution to establish the system's temperature:

(22)$$p( v_i) = \sqrt {\displaystyle{{m_i} \over {2\pi kT}}} \exp \left({-\displaystyle{{m_iv_i^2 } \over {2k_{\rm B}T}}} \right)$$

where k is Boltzmann's constant, T is temperature, m i and v i are the mass and velocity of atoms i, respectively. Because the ensemble generated by MD is a function of time, this technique is frequently used to assess the kinetics of intra- and intermolecular interactions, which we later describe in sections ‘Computational methods for predicting intramolecular dynamics of MAPIDs’ and ‘Computational methods for predicting the MAPID co-assembly’.

Similar to MC simulations, MD simulations yield trajectories of conformations, ${\vector X} = \{ {{\{ {\vector x} \} }_1, \;\;{\{ {\vector x} \} }_2, \;\ldots , \;\;{\{ {\vector x} \} }_n} \}$. MD simulations have been used to provide atomic resolution structural characterizations of IDRs in MAPIDs. For example, unbiased μs-long explicit MD simulations of the troponin complex with TnI IDRs were conducted to study its dynamics, stability, intramolecular interactions, and Ca2+ affinity (Cheng et al., Reference Cheng, Lindert, Kekenes-Huskey, Rao, Solaro, Rosevear, Amaro, McCulloch, McCammon and Regnier2014; Lindert et al., Reference Lindert, Cheng, Kekenes-Huskey, Regnier and McCammon2015; Zamora et al., Reference Zamora, Papadaki, Messer, Marston and Gould2016). The N-terminal IDR of TnI was shown to interact with TnC, whereupon the IDR's phosphorylation at S23/S24 reduces TnC Ca2+ affinity (Cheng et al., Reference Cheng, Lindert, Kekenes-Huskey, Rao, Solaro, Rosevear, Amaro, McCulloch, McCammon and Regnier2014; Zamora et al., Reference Zamora, Papadaki, Messer, Marston and Gould2016).

Brownian dynamics (BD) simulations BD facilitates coarsening of the system spatial resolution to lengthen the simulation time steps and ultimately the duration of the simulation. This is done by describing the influence of explicit solvent molecules on the solute with a random force and a dissipative friction term. In this way, the solute alone can be simulated without solvent, which permits much larger simulation step sizes:

(23)$$m{v}^{\prime} = {-}\nabla U( \vec{r}) -\gamma v + \sqrt {2\gamma k_{\rm B}Tn( t) } , \;\quad n( t) = {\cal N}( 0, \;1) $$
(24)$$v = {-}\nabla U( \vec{r}) /\gamma + \sqrt {2\gamma k_{\rm B}Tn( t) } /\gamma , \;\quad m{v}^{\prime}\to 0$$

where γ is a friction coefficient and ${\cal N}$ is a Gaussian distribution of random variates with a mean of zero and variance of 1. The latter equation represents an overdamped system, for which the inertial terms are insignificant. As with MD, BD generates a trajectory comprising protein conformations at successive time points. BD has been used to reveal the ensemble properties of several naturally occurring IDRs as well as engineered (Glu-Lys)25 IDRs (Ahn et al., Reference Ahn, Huber and McCammon2022). When coupled with an appropriate interaction potential, such as the coarse-grained force field for proteins (Ahn et al., Reference Ahn, Huber and McCammon2022) and a Debye–Huckle model (Eq. (9)), molecular properties such as inter-residue distances, the IDP's aggregation propensity, and ionic strength effects can be determined (Ahn et al., Reference Ahn, Huber and McCammon2022). BD has also been widely used to explore both the intramolecular and intermolecular dynamics of IDRs, as will be discussed in sections ‘Computational methods for predicting intramolecular dynamics of MAPIDs’ and ‘Computational methods for predicting the MAPID co-assembly’, respectively. This technique has been applied to the myofilament proteins TnC and Tm. Specifically, BD was used to estimate Ca2+ binding rates to TnC (Lindert et al., Reference Lindert, Kekenes-Huskey, Huber, Pierce and McCammon2012a) and to simulate the binding of Tm to actin (Aboelkassem et al., Reference Aboelkassem, McCabe, Huber, Regnier, McCammon and McCulloch2019).

Coarse-grained (CG) simulations CG represents another strategy of reducing a system's degrees of freedom to improve sampling efficiency. In general, CG schema treat each residue as a single bead located at the C$_\alpha$ atom, or two beads with the second bead representing a residue's side chain. One such CG scheme is the Martini model (Marrink et al., Reference Marrink, Risselada, Yefimov, Tieleman and de Vries2007) that provides an intermediate representation between CG and all-atom (AA). This is done primarily by representing large side chains by more than two beads. Compared to AA simulations, the degrees of freedom in CG are greatly reduced and thereby enable larger time steps than all-atom simulations.

CG molecular dynamics simulations have been used with IDPs (Ramis et al., Reference Ramis, Ortega-Castro, Casasnovas, Marino, Vilanova, Adrover and Frau2019), such as for investigating IDP-aggregation or liquid–liquid phase transition (LLPS) phenomena (Benayad et al., Reference Benayad, Blow, Stelzl and Hummer2021). LLPS describes the tendency of IDRs to form multi-valent PPIs with themselves or with other proteins. This aggregation can enable assemblies to phase separate from the surrounding solution (Harmon et al., Reference Harmon, Holehouse, Rosen and Pappu2017). It is believed that this phenomenon is used in biological systems to sequester biomolecules (Zhao and Zhang, Reference Zhao and Zhang2020). To model this behavior, lattice and CG simulations have been proposed for describing phases of multi-valent protein assemblies (Harmon et al., Reference Harmon, Holehouse, Rosen and Pappu2017). Recently, Sponga et al. showed that in the sarcomere, FATZ-1 contains IDRs that condense into a liquid phase, which may provide a new mechanism to interact with α-actinin (Sponga et al., Reference Sponga, Arolas, Schwarz, Jeffries, Chamorro, Kostan, Ghisleni, Drepper, Polyansky, Ribeiro, Pedron, Zawadzka-Kazimierczuk, Mlynek, Peterbauer, Doto, Schreiner, Hollerl, Mateos, Geist, Faulkner, Kozminski, Svergun, Warscheid, Zagrovic, Gautel, Konrat and Djinovi-Carugo2021). These models can benefit from coarse-graining the Hamiltonian to include only bond, Lennard-Jones, and bending potentials (Garaizar et al., Reference Garaizar, Sanchez-Burgos, Collepardo-Guevara and Espinosa2020). In parallel, new force fields are also being developed to better align these simulations with experiments (Regy et al., Reference Regy, Thompson, Kim and Mittal2021; Wohl et al., Reference Wohl, Jakubowski and Zheng2021).

Enhanced sampling techniques One of the most significant challenges in conducting simulations of IDRs arises from the proteins' conformational diversity and expansive timescales to match experimentally measured ensemble-level properties. This challenge stems from two limitations of MD methods: (1) force field accuracy and (2) sampling efficiency. A variety of force fields have been developed to improve IDR simulation accuracy, such as the ff14IDPSFF (Song et al., Reference Song, Luo and Chen2017). Experimental constraints are also frequently introduced to restrict the conformational space sampled during simulation, thus focusing the search on biologically relevant states (see reviews Bhattacharya and Lin, Reference Bhattacharya and Lin2019; Gong et al., Reference Gong, Zhang and Chen2021; Wang et al., Reference Wang, Grange, Wagner, Kho, Gautel and Raunser2021). Brute force, sub-ms length AAMD simulations have been performed to describe isolated IDP ensembles and binding mechanisms to partners (Robustelli et al., Reference Robustelli, Piana, Shaw and Shaw2020; Chang et al., Reference Chang, Wilson, Karunatilleke, Moselhy, Karttunen and Choy2021). However, these simulations are either performed on specialized machines such as Anton (Robustelli et al., Reference Robustelli, Piana and Shaw2018) or require extensive computational resources. A number of enhanced sampling techniques have been developed (Gong et al., Reference Gong, Zhang and Chen2021) to reduce the computational expense with conventional computing resources.

Temperature-based methods. Elevating a system's temperature increases the likelihood of the system to sample minima that are separated by large PES barriers (see Fig. 7). Replica exchange MD simulations exploit this principle by simultaneously running multiple replicas of the system at different target temperatures. During the simulations, states of neighboring replicas are exchanged at constant intervals, with the exchange probability defined as:

(25)$$p( \{ {\vector x}\} _i) \leftrightarrow ( \{ {\vector x}\} _j) = \min ( 1, \;\exp [ ( \displaystyle{1 \over {k_{\rm B}T_i}}-\displaystyle{1 \over {k_{\rm B}T_j}}) ( U( \{ {\vector x}\} _i) -U( \{ {\vector x}\} _j) ) ] ) $$

where T i and T j are the target temperatures, at which the replicas i and j are simulated. The number of replicas and the target temperatures are determined by the number of particles to set the exchange probability to roughly p ≈ 0.4 (Zhang et al., Reference Zhang, Wu and Duan2005). Temperature replica exchange MD has proven to be reliable in its many applications to both folded proteins and IDPs (Wu et al., Reference Wu, Weinstock, Narayanan, Levy and Baum2009; Wang et al., Reference Wang, Peng, Yu, Chen, Xu, Cai, Shao, Shi and Zhu2020). Many variations of replica exchange techniques, such as the Hamiltonian replica exchange MD (HREMD), replica exchange with hybrid tempering (Appadurai et al., Reference Appadurai, Nagesh and Srivastava2021), and Hamiltonian scaling in replica exchange with solute tempering (REST) (Liu et al., Reference Liu, Kim, Friesner and Berne2005) have been developed. For these variants, the main concept is to simulate only a subset of atoms to be at different temperatures, thus focusing sampling on the subsystem of interest. Applying these techniques can greatly shorten the simulation time for IDP studies. A recent example of this technique used unbiased all-atom molecular dynamics (AAMD) with HREMD to model the c-Src kinase N-terminal IDR ensemble over a 1 μs timescale (Shrestha et al., Reference Shrestha, Juneja, Zhang, Gurumoorthy, Borreguero, Urban, Cheng, Pingali, Smith, O’Neill and Petridis2019).

Fig. 7. Computational methods commonly used to model IDR structural properties. The IDR propensity of a structure can be predicted from its amino acid sequence by tens of established tools (Liu et al., Reference Liu, Wang and Liu2019), and structural states of IDP can be either explicitly modeled (Ozenne et al., Reference Ozenne, Bauer, Salmon, Huang, Jensen, Segard, Bernad, Charavay and Blackledge2012) or characterized by an implicit phase diagram (Holehouse et al., Reference Holehouse, Das, Ahad, Richardson and Pappu2017). Polymer models including the simplistic freely jointed monomer model and advanced models that account for intramonomer interactions can be used to characterize IDR ensembles (Milstein and Meiners, Reference Milstein, Meiners and Roberts2013; Schuler et al., Reference Schuler, Soranno, Hofmann and Nettels2016). Particle-based simulations are commonly used to predict IDR conformer structures and associated kinetics (Wang, Reference Wang2021). All atom simulations consider every individual atom in the system to determine detailed descriptions of the potential energy surface (PES) but are computationally intensive, while coarse-grained simulations lump atoms together to increase sampling efficiency at a modest loss of accuracy. Lastly, statistical models frequently use partition functions to obtain thermodynamic descriptions of IDRs (Hadzi et al., Reference Hadzi, Loris and Lah2021).

Modified potential. In complement to using elevated temperatures to cross PES barriers, reducing barrier heights by altering the potential can facilitate sampling. Biased potentials can be applied to user-defined collective variables (CVs) to enhance sampling. Adding boost potentials that raise the energy minima of a PES accelerates sampling by effectively lowering the energy barriers. Metadynamics is one such method that has been widely used to explore the conformational space of biomolecules and for determining free energy landscapes. In metadynamics, a biased potential with a Gaussian form is applied to a user-selected CV at a constant interval τ, which results in the potential (Laio and Parrinello, Reference Laio and Parrinello2002):

(26)$$V( \vec{s}, \;t) = \sum\limits_{n\tau } W ( n\tau < t) \exp \left({-\sum\limits_i^d {\displaystyle{{{( s_i-s_i^{( 0) } ( n\tau ) ) }^2} \over {2\sigma_i^2 }}} } \right)$$

where W() and σ i are the height and width of a bias potential, respectively. s i is the targeted state of the CV and $s_i^{( 0 ) } ( {n\tau } )$ is its instantaneous value at time . The advantage of metadynamics is that the sum of the added bias potentials over the simulation time course approximate the free energy profile along the CV:

(27)$$V( \vec{s}) = {-}G( \vec{s}) $$

In practice, however, the values of W, τ, and σ need to be carefully chosen for good performance.

The umbrella sampling method is a popular biased-sampling technique that enhances the sampling at windows along a CV by introducing a harmonic potential centered at each window's midpoint. The sampling of CVs along the windows is then used to recover the unbiased potential of mean force (PMF). A prominent challenge for the CV-based accelerating simulation technique is the choice of CVs. Choosing the appropriate CV that can capture a protein's biological-relevant degrees of freedom is challenging (Chen, Reference Chen2021). Bias-exchange metadynamics (BE-MTD) alleviates this problem by performing metadynamics simulations with different CVs to identify the optimal sampling path (Piana and Laio, Reference Piana and Laio2007). In BE-MTD, a metadynamics replica is performed for each CV, and the exchange of conformations from randomly paired replicas occurs at a constant time-interval with a Metropolis acceptance criterion (Eq. (13)). Lastly, machine learning techniques have also been trained to find optimal CVs (Chen, Reference Chen2021). For example, the autoencoder model, which contains an encoder to map high-dimensional data ($\{ {\vector x} \}$) to low-dimensional data ($\vec{s}$, CV) and a decoder to reverse the process, can be trained based on MD sampled configurations to identify an appropriate CV (Chen, Reference Chen2021).

Biased potentials can also be applied to broader sets of coordinates beyond CVs. Accelerated and Gaussian-accelerated MD are two related methods that are similar to metadynamics, but require no prior knowledge of CVs (Hamelberg et al., Reference Hamelberg, Mongan and McCammon2004; Miao et al., Reference Miao, Feher and McCammon2015). For Gaussian accelerated MD, the boost potential is applied system-wide to a protein's total potential, or the dihedral space (Miao et al., Reference Miao, Feher and McCammon2015), instead of user-selected CV, when the system's original potential is smaller than a threshold:

(28)$$U( {\vec{r}} ) = U_0( {\vec{r}} ) + \Delta U( {\vec{r}} ) , \;\quad {\rm if}\ U_0( {\vec{r}} ) \le E_{thres}$$

For this approach, the standard deviation of the boost potential $\sigma _{\Delta U( {\vec{r}} ) }$ is recommended to be less than 10k BT to ensure accurate reweighting for recovering the unbiased PES (Miao et al., Reference Miao, Feher and McCammon2015).

These categories of enhanced sampling techniques can be combined to improve predictions of an IDR ensemble. For example, in Li et al. (Reference Li, Casalini, Arosio and Salvalaglio2022), replica-exchange and metadynamics were combined to reveal ensemble properties of a 46 amino acid IDR derived from the DEAD-box protein DHH1. By using the advanced sampling techniques, the authors determined the free energy landscape as a function of secondary structure content and determined that the IDP has a low propensity for self-aggregation (Li et al., Reference Li, Casalini, Arosio and Salvalaglio2022). Lastly, while most of these strategies have been developed for explicit, AAMD simulations, in principle they can be adapted to any number of molecular simulation approaches.

Statistical mechanics models Statistical mechanics provides a framework to relate probabilistic representations of molecules to classical thermodynamics terms like the free energy. This is generally done by writing a partition function that enumerates the states available to a molecular system and their corresponding energies:

(29)$$\Omega ( \beta ) = \sum\limits_{x_i} {\exp } ( {-\beta {\cal H}( {x_1, \;x_2, \;\ldots } ) } ) , \;$$

where ${\cal H}$ represents an appropriate Hamiltonian for the system configuration (canonical) of interest. This partition function provides the appropriate normalization to determine the probability of a given state:

(30)$$P( {x_i} ) = \displaystyle{{\exp [ {-\beta {\cal H}( {x_i} ) } ] } \over {\Omega ( \beta ) }}$$

Standard thermodynamic relationships such as the free energy can then be defined with respect to the partition function:

(31)$$G = {-}k_BT\ln \vert \Omega \vert $$

As such, they are most frequently used for systems in equilibrium.

Hadzi et al. proposed a statistical thermodynamics model to describe the fuzzy ensemble of IDPs when bound to targets (Hadzi et al., Reference Hadzi, Loris and Lah2021). The core idea is that target-bound IDPs adopt heterogeneous structural states similar to the isolated state, but with higher fractions of α-helical content. Thus, in its bound state, an IDP's ensemble properties are governed by a coil-to-helix transition propensity, and constraints on this transition imposed by interactions with the target. In that coil–helix transition model, each residue in the polypeptide can adopt either a coil state or helix state. Additionally, a subset of these residues are ‘hot spots’ that interact with the target. This yields a partition function of the form:

(32)$$\Omega = \mathop \sum \limits_{2^{N_H}-1}^{h = 1} \Omega _{b, h}$$

where Ωb,h is a partition function for a set of hotspots (h) and N H is the total number of hotspots. Ωb,h in turn is based on a generating function that gives the statistical weight for a residue's propensity to undergo a coil-to-helix transition and a change in statistical weight based on its interaction energy, ΔG int, when bound with the target. These interaction energies can be obtained by MD simulations. This statistical model successfully predicted the changes of IDP helical content and binding affinity in two IDP–protein complexes driven by mutations in the IDPs, which were validated by NMR, CD, and isothermal titration calorimetry (ITC) binding assay data (Hadzi et al., Reference Hadzi, Loris and Lah2021).

Solvation models

Solvation is an important factor to consider when simulating an IDR ensemble, given that amino acid interactions with an aqueous solvent are strongly thermodynamically favorable for IDRs. The unstructured ensembles and dynamics of IDRs are more sensitive to protein/water interactions, compared to their folded counterparts (Wuttke et al., Reference Wuttke, Hofmann, Nettels, Borgia, Mittal, Best and Schuler2014). For this reason, current solvent models pose difficulties in accurately recapitulating experimentally determined IDR ensembles via simulation. This was recently illustrated in a study suggesting that the recently developed ‘general-purpose’ optimal opint charge (OPC) water model (Izadi et al., Reference Izadi, Anandakrishnan and Onufriev2014) still yields predictions of IDP ensembles that deviate from experiment, despite its improved performance over standard transferable intermolecular potential with 3 points (TIP3P) models (Shabane et al., Reference Shabane, Izadi and Onufriev2019). Efforts aiming to improve all-atom and CG solvation models for IDRs can be categorized into those improving the water model parameters alone or correcting the water/protein/ion interaction potentials. For example, refitting of the point charges in the TIP3P water model was reported to yield more accurately predicted IDP ensembles when validated against SAXS profiles (de Souza et al., Reference de Souza, Zariquiey and Bronowska2020). In addition, Gil Pineda et al. used a modified TIP3P water model combined with the CHARMM36m protein force field to correct a tendency for simulations to over-predict the compactness of IDPs (Gil Pineda et al., Reference Gil Pineda, Milko and He2020). Similarly, it was found that increasing the Martini forcefield's protein–solvent interaction potential greatly improved the agreement between sampled IDP ensembles with data from SAXS and NMR (Thomasen et al., Reference Thomasen, Pesce, Roesgaard, Tesei and Lindorff-Larsen2022). In tandem, optimization of metal ion parameters for different water models has been pursued toward reproducing experimentally determined hydration energies, coordination numbers, and ion-coordinated water exchange phenomena (Grotz and Schwierz, Reference Grotz and Schwierz2022). Lastly, there has been progress in adding energy terms to account for temperature-dependent protein stabilities into existing implicit solvent models to improve predictions for both IDP and globular proteins (Arsiccio et al., Reference Arsiccio, Pisano and Shea2022).

Multi-scale simulations

In contrast to ‘single-model’ approaches, other studies of IDRs have combined methodologies to enrich the diversity of sampling. As an example, a chain growth model has been combined with all-atom simulations to explore an IDP's ensemble (Pietrek et al., Reference Pietrek, Stelzl and Hummer2020). Pietrek et al. used AAMD simulations, with a local-to-global step-wise divide-and-conquer strategy to achieve comprehensive ensemble sampling of the α-synuclein IDP. This entailed first sampling the local fragments independently, assembling the fragments with a ‘chain-growth Monte Carlo’ strategy, and then simulating the intact IDP as a final step (Pietrek et al., Reference Pietrek, Stelzl and Hummer2020). This process has been improved by incorporating experimental chemical shift data into the fragment assembly (Stelzl et al., Reference Stelzl, Pietrek, Holla, Oroz, Sikora, Köfinger, Schuler, Zweckstetter and Hummer2022). Lastly, CG MD and AAMD can be combined, as shown by Garcia et al., to demonstrate the conformational space of a full-length IDP (Chvez-Garca et al., Reference Chvez-Garca, Hnin and Karttunen2022).

Bridging computer models with experiments

Simulations in section ‘Computational approaches’ are parameterized, validated, and refined using the experimental techniques listed in section ‘Experimental techniques’. We now discuss how comparisons between experimental data are made and how experimental data are used to constrain searches.

A variety of strategies have been proposed for aligning modeling results with, and validation against, experimental data. In principle, experimental data (such as FRET efficiency) and computational data (such as residue–residue distance) should yield similar distances between probed amino acids. However, in practice there may be limited overlap because of the intrinsically different spatiotemporal scales, at which these data are collected (Best, Reference Best2017) and measurement uncertainties. However, there are strategies that mitigate these comparisons, such as for co-aligning with NMR (order parameters (Lipari et al., Reference Lipari, Szabo and Levy1982)), SAXS (radius of gyration (Henriques et al., Reference Henriques, Cragnell and Skepo2015)), and fluorescence (pairwise distance (Metskas and Rhoades, Reference Metskas and Rhoades2015)). One such example includes a study that combined SAXS and NMR with the structure generation method Flexible-Meccano to characterize the ensemble of an IDP from the Sendai virus (Bernadó et al., Reference Bernadó, Blanchard, Timmins, Marion and Ruigrok2005). In that study, the conformations of the IDP were iteratively generated via the ‘Flexible-Meccano’ modeling method, and the RDC and SAXS data predicted from these conformations were then compared against experimental values until convergence (Bernadó et al., Reference Bernadó, Blanchard, Timmins, Marion and Ruigrok2005).

Similar approaches have been applied to the MAPIDs TnI, MyBPC3, and myotilin. Metskas et al. studied the IDRs in TnI, using MD simulations to probe the distances between multiple positions of TnI relative to TnC. These distances were compared against experimental FRET data (Metskas and Rhoades, Reference Metskas and Rhoades2015). Since the FRET observable time resolution (~1 ms) was orders of magnitude longer than the protein's conformational sampling (ns to μs), the experimental observable was reflective of the protein's average dynamics. Toward that end, the authors used short MD replicas in parallel, starting from different initial structures to capture the protein's conformational dynamics for direct comparison against FRET efficiency (Metskas and Rhoades, Reference Metskas and Rhoades2015). Michie et al. (Reference Michie, Kwan, Tung, Guss and Trewhella2016) used a similar strategy to identify a linker in the M-domain of MyBPC3 for binding calmodulin via SAXS scattering and computational modeling. Another excellent MAPID study that utilized both experiment and simulation was performed by Kostan et al. (Reference Kostan, Pavšič, Puž, Schwarz, Drepper, Molt, Graewert, Schreiner, Sajko, Ven, Onipe, Svergun, Warscheid, Konrat, Fürst, Lenarčič and Djinović-Carugo2021). Conformations comprising myotilin's ensemble were selected from a pool of structures generated by the ‘EOM’ approach (Bernadó et al., Reference Bernadó, Mylonas, Petoukhov, Blackledge and Svergun2007) to match experimental SAXS profiles. These conformations, together with biochemical binding assays, provide an integrative structural model that rationalizes the mechanism, by which myotilin's IDR binds to F-actin (Kostan et al., Reference Kostan, Pavšič, Puž, Schwarz, Drepper, Molt, Graewert, Schreiner, Sajko, Ven, Onipe, Svergun, Warscheid, Konrat, Fürst, Lenarčič and Djinović-Carugo2021).

Other approaches have used Bayesian inference to relate predicted structural models to experimental data. With this framework, the conditional probability of simulating a protein conformation y x, given an experimental observation, θ Rg, is determined from Bayes theorem. This observation, for instance, could be a value of R g implied from SAXS data. This probability, P(y x|θ Rg), is known as the posterior distribution. The posterior is determined from the prior distribution, P(y x), and the likelihood, P(θ Rg|y x). The prior distribution is the probability that the simulated protein yields a conformation y x. The likelihood is the probability of making an experimental observation θ Rg, given the simulated conformation y x. These relationships culminate in the equalities:

(33)$$P( {y_x\vert \theta_{R{\rm g}}} ) = \displaystyle{{P( {\theta_{R{\rm g}}\vert y_x} ) P( {y_x} ) } \over {P( {\theta_{R{\rm g}}} ) }}$$
(34)$$ = \displaystyle{{P( {\theta_{R{\rm g}}\vert y_x} ) P( {y_x} ) } \over {\int_y P ( {\theta_{R{\rm g}}\vert y_x} ) P( y_x) dy}}$$

The denominator in Eq. (33) represents the probability of experimentally observing the value θ Rg, and can be estimated from Eq. (34).

A study from Fisher et al. proposed a variation of this approach to determine population weights ($\overrightarrow {w( y ) }$) of simulated protein structures that are consistent with a set of experimental measurements ($\vec{\theta }$), using a posterior $P( \overrightarrow {w( y ) } \vert \vec{\theta })$. The population weights were set by the Boltzmann factor (Eq. (30)), using energies calculated for all conformations. The prior $P( {\vec{w}} )$ was based on Gaussian distributions using transformed representations of the weights. Importantly, the likelihood used the expected value of an experimental observable like R g from the weighted simulated structures, ${\cal E}( \theta _i\vert \vec{w})$, as well as the error in the predicted observable, $\epsilon _{R{\rm g}}$:

(35)$$P( \theta _i\vert \vec{w}) = \displaystyle{1 \over {\sqrt {2\pi \epsilon _{R{\rm g}}} }}\exp \left[{-\displaystyle{{{( \theta_i-{\cal E}( \theta_i\vert \vec{w}) ) }^2} \over {2\epsilon_{R{\rm g}}}}} \right]$$

In practice, the authors generated an ensemble of protein conformations via MD simulations, after which their Bayesian framework was used to determine the most likely conformation weights that agreed with RDCs or chemical shifts from NMR (Fisher et al., Reference Fisher, Huang and Stultz2010). Another variation of this approach defined an error function for SAXS data to penalize overfitting due to extraneous degrees of freedom (Bowerman et al., Reference Bowerman, Curtis, Clayton, Brookes and Wereszczynski2019). An example application to MAPIDs includes a study of MyBPC3, in which the Bayes inference is used to align simulation results with SAXS and NMR data. This strategy revealed that the flexible linker connecting the M-domain and C2 domain allows the two domains to sample diverse interdomain orientations, thus imparting considerable flexibility to the construct (Potrzebowski et al., Reference Potrzebowski, Trewhella and Andre2018).

Computational methods for predicting intramolecular dynamics of MAPIDs

Problem and application

The kinetics with which conformations exchange with one another are equally important relative to the breadth of conformations adopted by an IDP, e.g.

$$\{ {\vector x} \} _1\mathrel{\mathop{\kern0pt\rightleftharpoons}\limits_{{\rm k}_{ 21}}^{{\rm k}_{ 12}}} \{ {\vector x} \} _ 2$$

where $\{ {\vector x} \} _i$ refers to the conformation i. This information is critical for understanding the timescale of MAPID functions relative to other physiological phenomena.

As an example, the TnI switch peptide is disordered at resting (diastolic) Ca2+ levels. Binding of this peptide to TnC's hydrophobic patch promotes a disorder-to-order transition of the peptide to form a folded α-helix in response to elevated cytosolic Ca2+ (Metskas and Rhoades, Reference Metskas and Rhoades2016). The timescale for this conformational search and fold process must occur within the brief rise in calcium that accompanies a typical heartbeat. For instance, the breadth of the TnI C-terminal IDR ensemble determines its effective concentration for steady-state measurements (section ‘Implicit/semi-analytic representations’). However, it is the rate at which the IDR samples the appropriate position and conformation to bind TnC, relative to the duration of elevated calcium, that will determine the rate of activating the thin filament. Similarly, the intrinsic timescales for myosin's dynamic IDR loops to interact with actin could impact the rate of cross-bridge formation (Gurel et al., Reference Gurel, Kim, Ruijgrok, Omabegho, Bryant and Alushin2017; Doran et al., Reference Doran, Pavadai, Rynkiewicz, Walklate, Bullitt, Moore, Regnier, Geeves and Lehman2020). Altogether, the dynamics of these IDRs and those of other MAPIDs collectively influence the kinetics of force generation (Fig. 8).

Fig. 8. (a) Dynamics of an IDP ensemble represented by a Markov state model (MSM). (b) Experimental methods for structure determination and their temporal resolutions. Timescale information of NMR, X-ray, cryo-EM, and SAXS are taken from Ban (Reference Ban and Miller2020). FRET time resolution spans ns to seconds (Okamoto and Sako, Reference Okamoto and Sako2017). Size information: EM is appropriate for proteins >100 kD (Yeates et al., Reference Yeates, Agdanowski and Liu2020). NMR is most suitable for <30 kD proteins (Xu et al., Reference Xu, Zheng, Fan and Yang2006). X-ray can solve structures up to 4000 kD (PDB website statistics). FRET is used for medium-sized proteins but has been reported for up to a 540 kD protein (Sielaff et al., Reference Sielaff, Dienerowitz and Dienerowitz2022).

Experimental techniques

NMR

Probes of IDR conformational dynamics can be resolved via NMR experiments through analyzing spin relaxation data (Ban et al., Reference Ban, Smith, Groot, Griesinger and Lee2017). For instance, residue mobilities and interactions can be obtained by fitting longitudinal (R1)/transverse (R2) relaxation rates and heteronuclear steady-state NOEs (Sibille and Bernado, Reference Sibille and Bernado2012). Such techniques provide high structural resolution, although timescales can be limited to μs-ms for R1/R2 relaxations and ps-ns for NOEs (Palmer, Reference Palmer2004). RDCs can also be used to probe protein dynamics at atomic resolution in the ps to ms ranges (Tolman and Ruan, Reference Tolman and Ruan2006). As an example, N15 relaxation provides ps to ns time resolution (Hwang et al., Reference Hwang, Cai, Pineda-Sanabria, Corson and Sykes2014), though interconversion of IDP conformations may approach timescales of μs or longer (Bernetti et al., Reference Bernetti, Masetti, Pietrucci, Blackledge, Jensen, Recanatini, Mollica and Cavalli2017). Abyzov et al. also demonstrated that temperature-dependent NMR can yield ‘local’ activation energies for dynamic modes of an IDR ensemble (Abyzov et al., Reference Abyzov, Salvi, Schneider, Maurin, Ruigrok, Jensen and Blackledge2016). NMR relaxation techniques have also been used with TnI. Hwang et al. as an example assessed the dynamics of TnI's N-terminus (residues M1–K37) and found that it remains intrinsically disordered even after binding to TnC (Hwang et al., Reference Hwang, Cai, Pineda-Sanabria, Corson and Sykes2014). In addition, relaxation rate constants and NOEs collected for MyBPC3 indicated that the linker spanning its tri-helix bundle and C2 domain exchanges conformations on a ps to μs timescale (Michie et al., Reference Michie, Kwan, Tung, Guss and Trewhella2016), that is much faster than the typical heart beat.

Time-resolved X-ray and SAXS

Time-resolved X-ray crystallography enables observations of how electron density maps of protein domains evolve in time, which can provide functional insights (Schotte et al., Reference Schotte, Lim, Jackson, Smirnov, Soman, Olson, Phillips, Wulff and Anfinrud2003). While applications to MAPIDs have not yet been reported in the literature, the technique was used with myoglobin to reveal a series of structural intermediates exchanging on a 150 ps timescale during its functional cycle (Schotte et al., Reference Schotte, Lim, Jackson, Smirnov, Soman, Olson, Phillips, Wulff and Anfinrud2003). Time-resolved X-ray scattering was also applied to resolve quaternary structure dynamics on a 316 ns to 100 μs range for carboxyhemoglobin (Cho et al., Reference Cho, Schotte, Stadnytskyi and and Anfinrud2021). In addition, SAXS has been used to complement NMR studies. In this context, SAXS can provide the general shape of an IDR ensemble, while NMR is used to reveal structural dynamics at ps–ns, μs, and ms timescales (Sibille and Bernado, Reference Sibille and Bernado2012; Schneider et al., Reference Schneider, Blackledge and Jensen2019). For example, the combination of SAXS and NMR enabled both conformational and dynamic characterization of IDRs in the ribosomal L12 protein (Sibille and Bernado, Reference Sibille and Bernado2012), for which not only the ensemble of structures, but also their conformational transition timescales are obtained.

Fluorescence techniques

Fluorescence-based techniques such as time-resolved fluorescence anisotropy are powerful tools for resolving protein motions occurring at different timescales. Time-resolved fluorescence anisotropy is well-suited for monitoring local backbone rotation and long-range correlation kinetics (Das et al., Reference Das, Arora and Mukhopadhyay2021). One example from Grupi and Haas (Reference Grupi and Haas2011) introduced fluorescent labels at different positions of the α-synuclein IDP. Time-resolved FRET was then used to determine the α-synuclein end-to-end distance distributions, which revealed the disordered nature of the protein, as well as its rapid intrachain diffusion dynamics (Grupi and Haas Reference Grupi and Haas2011). Another example revealed TnT's flexible linker conformations on the full cardiac thin filament. The time-resolved FRET data measured from labels in the IDR linker were then used as constraints to bias MD sampling of the TnT structure (Deranek et al., Reference Deranek, Baldo, Lynn, Schwartz and Tardiff2022). Fluorescence correlation spectroscopy is another technique that can be used to probe conformational changes occurring over 20–300 ns (Lee et al., Reference Lee, Moran-Gutierrez and Deniz2015), but applications to myofilament proteins appear to be limited to determining protein concentrations in vivo, such as by Duggal et al. (Reference Duggal, Requena, Nagwekar, Raut, Rich, Das, Patel, Gryczynski, Fudala, Gryczynski, Blair, Campbell and Borejdo2017).

Computational approaches

Implicit, semi-analytical approaches

Examples of implicit, semi-analytical methods for intramolecular IDP kinetics are fewer in number relative to the structural models introduced in the previous section. One model of note gives the rate, k, of two ends of a polymer to contact one another (Grupi and Haas, Reference Grupi and Haas2011):

(36)$$\displaystyle{1 \over k} = \displaystyle{{R^2} \over {3D}}\left({\displaystyle{{\sqrt \pi } \over {2\alpha }} + \ln 2-1-\displaystyle{{\sqrt \pi } \over 2}\alpha + \displaystyle{4 \over 3}\alpha^2} \right)$$

where D is the intramolecular end-to-end diffusion coefficient, $R={\sqrt {\langle r^2\rangle}}$ is the root mean square distance between the interacting residues, a is the radius of the residue, α = (3/2)0.5(a/R), and D is the intramolecular diffusion coefficient. Similarly, there have been efforts to determine the kinetics of forming interior loops in polypeptide chains. For instance, loop closure kinetics for a WLC can be estimated if the PMFs are known in advance (Hyeon and Thirumalai, Reference Hyeon and Thirumalai2006).

Explicit simulations of intramolecular IDR kinetics

Explicit models of IDP kinetics are common. Early approaches relied on explicit definitions of reaction coordinates (RCs), which are coordinates that describe the progress of a system exchanging between two states, such as dynamic changes in end-to-end distances (Doucet et al., Reference Doucet, Roitberg and Hagen2007). Relating these dynamic changes to kinetics has been done using approaches based on the Smoluchowski equation (Hamelberg et al., Reference Hamelberg, Shen and Andrew McCammon2005) to estimate dynamics along a one-dimensional PMF (Szabo et al., Reference Szabo, Schulten and Schulten1980; Hamelberg et al., Reference Hamelberg, Shen and Andrew McCammon2005):

(37)$$\displaystyle{{\partial p( x, \;t) } \over {\partial t}} = \displaystyle{\partial \over {\partial x}}\left[{D( x) {\exp }^{-\beta U( x) }\displaystyle{\partial \over {\partial x}}( {{\exp }^{\beta U( x) }p} ) } \right]$$

where p(x, t) is the time-dependent probability density along the RC, D is the diffusion coefficient, and U(x) is the PMF. Other approaches include using Kramer's rule (Berezhkovskii and Szabo, Reference Berezhkovskii and Szabo2005) to calculate the transition rate (Berezhkovskii and Szabo, Reference Berezhkovskii and Szabo2005):

(38)$$k = \left[{\left({\int_{x_b-\Delta x}^{x_b + \Delta x} {{\exp }^{-\beta U( x) }} dx} \right)\left({\int_{x_T-\Delta x}^{x_T + \Delta x} {\displaystyle{{{\exp }^{\beta U( x) }} \over {D( x) }}} dx} \right)} \right]^{{-}1}$$

where x is the RC, while x b and x T represent the initial state and transition state, respectively. Hyeon et al. also used Kramer's rule to predict passage times over potentials of mean force based on the equilibrium distributions of the chains (Hyeon and Thirumalai, Reference Hyeon and Thirumalai2006).

These approaches rely on estimates of D(s), which can be determined via time-resolved FRET (Grupi and Haas, Reference Grupi and Haas2011), or by using methods we discuss below. In many cases, a single (spatially uniform) diffusion coefficient suffices. This can be obtained via experiment or simulation by estimating a particle's mean square displacements, ⟨x 2⟩, which are related to the diffusion coefficient in one dimension:

(39)$$\mathop {\lim }\limits_{t\to \infty } \langle ( x( t) -x( 0) ) ^2\rangle = 2Dt$$
(40)$$\Rightarrow \langle x^2\rangle = 2Dt$$

The degree of disorder in an IDR can also be used to refine this constant (Chu and Wang, Reference Chu and Wang2019). Other situations, especially those that use D(x) in parallel with potentials of mean force, benefit from estimates that are spatially dependent. In these cases, autocorrelation functions of a solute's positions or velocities from explicit simulations can be helpful to obtain spatially resolved diffusion coefficients (Hummer, Reference Hummer2005) along a PMF and its barriers. Here one can rely on the principle that the autocorrelation function for Markovian processes decays exponentially over time with a correlation time τ c:

(41)$$\langle x( 0) x( t) \rangle = \langle x^2\rangle \exp ( {-}t/\tau _{\rm c}) $$
(42)$$\tau _c = \displaystyle{1 \over {\langle x^2\rangle }}\int_0^\infty {\langle x( } 0) x( t) \rangle dt$$

The correlation time is related to the particle's friction coefficient, ζ, and mass via τ c = m/ζ. The Einstein relation D = k BT/ζ then yields the diffusion coefficient from ζ. A derivation using harmonic forces to constrain the system along the PMF is provided in the appendix of Pace et al. (Reference Pace, Rahmaninejad, Sun and Kekenes-Huskey2021).

A key concern with such approaches is that force fields are generally optimized for equilibrium structures, not their dynamics, therefore the kinetic barriers to conformational motions may be poorly described (Ponder and Case, Reference Ponder and Case2003). In addition, states with large energy barriers are much less frequently sampled or even inaccessible for conventional MD simulations, which could pose difficulties in identifying the most probable RCs. In general, RCs can be difficult to define or may consist of many parallel pathways (Hinczewski et al., Reference Hinczewski, Hansen, Dzubiella and Netz2010).

To partially address these limitations, enhanced sampling techniques have been used to sample regions that are separated by large energy barriers. For instance, accelerated MD simulations were reweighted and used Kramer's rule (Eq. (38)) to estimate the unbiased kinetics of intramolecular transitions between conformations (Hamelberg et al., Reference Hamelberg, Shen and Andrew McCammon2005). More recently, in Bernetti et al. (Reference Bernetti, Masetti, Pietrucci, Blackledge, Jensen, Recanatini, Mollica and Cavalli2017) metadynamics was used to reveal the PES of an IDP. Using a post-processing analysis called ‘Bin-Based Kinetic Model’, rate constants were recovered from the biased metadynamics trajectories.

Markov techniques A shortcoming of the preceding approaches is that an RC is needed to be defined a priori. Markov state modeling techniques (such as msmbuilder (Harrigan et al., Reference Harrigan, Sultan, Hernández, Husic, Eastman, Schwantes, Beauchamp, McGibbon and Pande2017) and pyemma (Scherer et al., Reference Scherer, Trendelkamp-Schroer, Paul, Pérez-Hernández, Hoffmann, Plattner, Wehmeyer, Prinz and Noé2015)) bridge this gap by allowing the RCs to be directly determined from the MD simulation data. In brief, long simulations are performed to yield trajectories that span the conformational space of the protein. Then, the conformation ensembles from these simulations are discretized into microstates, based on a user-defined metric like RMSD (Prinz et al., Reference Prinz, Wu, Sarich, Keller, Senne, Held, Chodera, Schtte and No2011; Husic and Pande, Reference Husic and Pande2018), from which rate constants for exchanging between microstates are determined. This approach is based on the idea that the time-dependent change in the probability of a given state can be described as a rate of transition out of (or into) the state:

(43)$$\displaystyle{{dp_j} \over {dt}} = {-}p_j\sum\limits_{i\ne j} {k_{\,ji}} + \sum\limits_{i\ne j} {k_{ij}p_i} $$

or more generally, for multiple states:

(44)$$\displaystyle{{d{\vector p}( {\vector t}) } \over {dt}} = {\vector K}( {\delta t} ) {\vector p}( t) $$

The relationship:

(45)$${\vector p}( t) = \exp ( {\vector K}t) {\vector p}_0$$

can then be used to determine the state probabilities at time t.

If ${\vector K}$ from Eq. (44) is not known, which is generally the case for Markov state model (MSM) approaches, the change in probabilities can be expressed in terms of a transition probability matrix, ${\vector T}$. This matrix evolves the state probabilities by some δt:

(46)$${\vector p}( {t + \delta t} ) = {\vector T}( {\delta t} ) {\vector p}( t ) $$

The transition matrix is populated by counting the transition events from microstate i to j occurring within δt (the lag time), as calculated from the discretized trajectory. If the model is Markovian for n consecutive lag time periods, the following holds:

(47)$${\vector p}( t + n\delta t) = {\vector T}^n( \delta t) {\vector p}( t) ; \;$$

that is, as n → ∞, the steady-state distribution (${\vector p}_0$) is obtained. For systems that behave as Markovian processes (namely the transition probability is only dependent on the current state and is independent of previous states), the implied timescale of the m th eigenvalue, t m becomes independent of the lag time δt:

(48)$$t_m = {-}\displaystyle{{\delta t} \over {\ln \lambda _m}}$$

where λ m is the m th eigenvalue of the transition probability matrix ${\vector T}$. In practice, this implied timescale t m is used to validate that the model built from simulations is Markovian. Transition matrices from this approach can then be used to find ${\vector K}$ by recognizing that (Polizzi et al., Reference Polizzi, Therien and Beratan2016):

(49)$${\vector K} = \mathop {\lim }\limits_{\delta t\to 0} \left({\displaystyle{{{\vector T}-{\vector I}} \over {\delta t}}} \right)$$

Once the transition matrix is determined, useful quantities such as the mean first passage time (MFPT), ⟨t fps can be estimated. The MFPT represents the average time for a system to transition into an absorbing state, s. A germane example of an MFPT would be to estimate the folding rate of a protein, where the sth state is the folded end-point and all other states are unfolded intermediates (Dai et al., Reference Dai, Sengupta and Levy2015). This quantity can be determined by:

(50)$$\langle t_{\,fp}\rangle _s = \sum\limits_{i\ne s} {r_i} $$

where r i is the residence time of state i. The residence times for all states can be determined from the rate matrix ${\vector K}$, following the derivation from Polizzi et al. (Reference Polizzi, Therien and Beratan2016). In their approach, a rate matrix, $\widetilde{{\vector K}}$, is defined that includes an absorbing condition for state s (e.g. $k_{si} = 0\ \forall i\ne s$). This allows the residence times in each state to be determined by:

(51)$${\vector r} = \int_0^\infty {\exp } ( \widetilde{{{\vector K}t}}) \widetilde{{{\vector p}_0}}dt$$
(52)$$\Rightarrow r_i = [ {-{\widetilde{{\vector K}}}^{{-}1}\widetilde{{{\vector p}_0}}} ] _i$$

where $\widetilde{{{\vector p}_0}}$ is the probability distribution with p s = 0.

The thermodynamics and transition kinetics between the metastable states can then be readily obtained from the MSM model (Qiao et al., Reference Qiao, Bowman and Huang2013). The computational complexity of these simulations is daunting as increasingly larger numbers of states are considered; in general, dimension reduction techniques (e.g. time-lagged independent component analysis (Pérez-Hernández et al., Reference Pérez-Hernández, Paul, Giorgino, De Fabritiis and Noé2013)) and state grouping methods (e.g. k-means clustering) are needed to achieve low-dimensional data and a manageable number of states (Scherer et al., Reference Scherer, Trendelkamp-Schroer, Paul, Pérez-Hernández, Hoffmann, Plattner, Wehmeyer, Prinz and Noé2015).

Brownian dynamics (BD) Brownian dynamics that describe molecular motions of systems with user-interested potentials could provide unique insights into IDP dynamics. When compared to polymer models like the WLC formalism, BD can not only provide ensemble properties like compactness, but also kinetic information such as contact rates, as done by Mühle et al. (Reference Mühle, Zhou, Ghosh and Enderlein2019). In their study, the dynamics of IDPs with different chain lengths were simulated with a minimal BD model, in which IDPs are treated as beads connected by fixed bonds. The authors showed that hydrodynamic interactions and excluded volume effects were key factors that governed the IDP's dynamics, by comparing experimental measurements of end-to-end contacts obtained from fluorescence correlation data (Mühle et al., Reference Mühle, Zhou, Ghosh and Enderlein2019). The dynamic properties of higher-order structures, such as those formed by the co-assembly of the conformations of intrinsically disordered Phe-Gly repeats, can also be probed by BD (Moussavi-Baygi and Mofrad, Reference Moussavi-Baygi and Mofrad2016). The fast recovery of high-order structures bestows IDPs with a tolerance to perturbations that may protect their function (Moussavi-Baygi and Mofrad, Reference Moussavi-Baygi and Mofrad2016).

Multi-scale approaches

Multiscale methods can also be used to estimate kinetics. One such example relies on both state discretization (like MSM) and diffusion along a PES (like the Smoluchowski model). This combination was used to reveal the intramolecular rate constants for an IDP of the Sendai virus nucleoprotein (Bernetti et al., Reference Bernetti, Masetti, Pietrucci, Blackledge, Jensen, Recanatini, Mollica and Cavalli2017). In that study, the PES of the IDP was first obtained by metadynamics simulations. The PES was then used to guide the state discretization to achieve a tractable number of states (Bernetti et al., Reference Bernetti, Masetti, Pietrucci, Blackledge, Jensen, Recanatini, Mollica and Cavalli2017). The transition rate between two metastable states a to b was then given by:

(53)$$k_{ij} = k_{ij}^0 \exp \left({\displaystyle{{G_i-G_j} \over {2k_BT}}} \right)$$

where G i − G j is the free energy difference between two minima and $k_{ij}^0$ is the transition rate on a flat PES. Here, k 0 is a function of a diffusion coefficient D and the barrier height (Eq. (37)). Multiple MD replica simulations were then performed to sample transition events between discretized states to estimate the transition probability matrix (see Eq. (48)). Lastly, kinetic MC simulations based on the transition probability matrix and trial values of D as inputs were performed. The kinetic MC model parameterization that recovered the observed state from the MD replicas was used for determining the appropriate D (Bernetti et al., Reference Bernetti, Masetti, Pietrucci, Blackledge, Jensen, Recanatini, Mollica and Cavalli2017). Combinations of the computational techniques discussed in this section that have been applied to MAPIDs are summarized in Table 1.

Computational methods for predicting the MAPID co-assembly

Problem and application

Building on the previous sections where the structure or dynamics of isolated MAPIDs are considered, in this section we seek to simulate how intrinsic disorder determines the structures of associated proteins and their rates of binding. This is important for probing how protein structures switch between unfolded and folded conformations that are spontaneous (conformational selection) or induced during binding (induced-fit) (Sugase et al., Reference Sugase, Dyson and Wright2007). Similarly, we can learn how proteins that host many SLIMs within a dynamic sequence can utilize fast on/off binding rates to facilitate high-affinity interactions (Hough et al., Reference Hough, Dutta, Sparks, Temel, Kamal, Tetenbaum-Novatt, Rout and Cowburn2015). These introduce two main challenges: (1) How to computationally calculate the transition kinetics between biologically relevant states of an IDP and (2) How to relate these kinetics to IDP functions.

IDPs structures and their association kinetics are particularly relevant to myofilament contraction and its dynamic regulation. The binding of TnI's C-terminal domain with Ca2+-saturated TnC is a classical example, whereby the disordered TnI switch peptide undergoes an unfolded to folded transition when bound. Simulating these interactions still needs theoretical developments that can more accurately capture the structures and dynamics of these interactions (Schuler et al., Reference Schuler, Borgia, Borgia, Heidarsson, Holmstrom, Nettels and Sottini2020).

Experimental techniques

Binding assays

Binding assays are frequently used to assess the association of IDRs with other protein targets. Electrophoresis (see section ‘Experimental techniques’) is routinely used to determine the extent, to which two proteins form a complex, based on differences in the migration of the complex versus the isolated components. This assay for instance was used to assess the binding of tropomyosin to tropomodulin (Greenfield et al., Reference Greenfield, Kostyukova and Hitchcock-DeGregori2005; Kostyukova et al., Reference Kostyukova, Hitchcock-DeGregori and Greenfield2007).

F-actin co-sedimentation is another established method to investigate the direct interaction of proteins with the actin filament (Srivastava and Barber, Reference Srivastava and Barber2008). Co-sedimentation consists of two steps: (1) incubation of purified proteins with actin, (2) centrifugation to pellet actin and analysis of the proteins that co-sediment with actin (Srivastava and Barber, Reference Srivastava and Barber2008). Co-sedimentation was also used to reveal that PEVK motifs in titin bind to actin in a Ca2+-dependent manner (Linke et al., Reference Linke, Kulke, Li, Fujita-Becker, Neagoe, Manstein, Gautel and Fernandez2002).

Isothermal titration calorimetry (ITC) is a very accurate and commonly used technique capable of measuring the energetics of biomolecules binding over a wide range of affinities (10−3 to 10−12 M−1) (Velázquez-Campoy et al., Reference Velázquez-Campoy, Ohtaka, Nezami, Muzammil and Freire2004). This method works by incrementally titrating in a reagent in excess of its binding partner. The heat released from the association event is exchanged with a bath and measured. Analyzing the ITC curves allows the dissociation constant, stoichiometry, enthalpy, and entropy to be determined simultaneously (Velázquez-Campoy et al., Reference Velázquez-Campoy, Ohtaka, Nezami, Muzammil and Freire2004). ITC was used to determine the binding affinity of ankyrin's auto-inhibitory IDR (Chen et al., Reference Chen, Li, Wang, Wei and Zhang2017). These techniques unfortunately do not provide structural and kinetic information about the IDP binding process.

Fluorescence spectroscopy

The fluorescence techniques introduced in section ‘Experimental techniques’ lend themselves to probing the binding between IDRs and their targets (Lee et al., Reference Lee, Moran-Gutierrez and Deniz2015). As one example, time-resolved FRET has been used in tandem with MD simulations to identify functionally important conformations of TnT's IDR linker when bound to the thin filament (Deranek et al., Reference Deranek, Baldo, Lynn, Schwartz and Tardiff2022). Another example monitored differences in intrinsic fluorescence as a probe for conformational changes during TnT/TnC binding; this study also revealed that a DCM-associated mutation in TnC enhanced their binding (Johnston et al., Reference Johnston, Landim-Vieira, Marques, Oliveira, Gonzalez-Martinez, Moraes, He, Iqbal, Wilnai, Birk, Zucker, Silva, Chase and Pinto2019). Intrinsic fluorescence has also been used to probe the interaction between titin's PEVK motifs and actin in the presence of S100A1 and Ca2+ (Yamasaki et al., Reference Yamasaki, Berri, Wu, Trombitas, McNabb, Kellermayer, Witt, Labeit, Labeit, Greaser and Granzier2001). The kinetics of assembly are additionally amenable to stopped-flow studies that monitor changes in intrinsic fluorescence following the rapid mixing of two species (Zheng et al., Reference Zheng, Bi, Li, Podariu and Hage2015). The time-dependent changes in intrinsic fluorescence can then be fit to rate laws to determine association kinetics (Zheng et al., Reference Zheng, Bi, Li, Podariu and Hage2015). Lastly, in recent years, ‘BRET’ techniques that utilize bioluminescence, such as from a luciferase protein donor, to facilitate FRET between a donor/acceptor pair have gained momentum toward characterizing PPIs in vivo (De et al., Reference De, Jasani, Arora and Gambhir2013).

NMR

NMR techniques are also heavily utilized for monitoring the kinetics and structures of IDR/protein association. Among these, chemical shifts are the most frequently used to monitor the assembly of MAPID structures. As an example, the binding of tropomodulin's N-terminal IDR to tropomyosin was studied by collecting 15N-1H HSQC spectra (Greenfield et al., Reference Greenfield, Kostyukova and Hitchcock-DeGregori2005). The IDR residues exhibit altered chemical shifts upon addition of tropomyosin, therefore allowing the determination of binding sites within the protein's IDR (Greenfield et al., Reference Greenfield, Kostyukova and Hitchcock-DeGregori2005). Similarly, Hwang et al. utilized multidimensional solution NMR spectroscopy to measure complex formation between a TnI fragment (residues M1–G73) and intact TnC. This study indicates that the TnI fragment gains helical content after binding to the C-domain of TnC (Hwang et al., Reference Hwang, Cai, Pineda-Sanabria, Corson and Sykes2014).

Other techniques

Other techniques can provide data for probing putative interaction sites between binding species. Mass spectrometry (MS) as an example is an increasingly used technique to determine the binding of IDR-containing species and, in some cases, the amino acids forming the protein/protein interface. Cross-linking mass spectrometry is a popular approach for the latter by identifying adjacent amino acids bridging PPI interfaces (Merkley et al., Reference Merkley, Cort and Adkins2013). Specifically, characterizing chemically cross-linked peptide fragments by mass spectrometry provides residue–residue interaction information. This method has been applied to TnT's C-terminal IDR binding to TnC, where the direct binding of residues N281–K286 from TnT to TnC residues M1–K6 was determined (Johnston et al., Reference Johnston, Landim-Vieira, Marques, Oliveira, Gonzalez-Martinez, Moraes, He, Iqbal, Wilnai, Birk, Zucker, Silva, Chase and Pinto2019).

Another MS-based technique, HDXMS, is frequently used to probe protein complexes. With HDXMS, protons on buried amino acids undergo less frequent deuterium exchange compared to solvent-exposed sites. Differences in deuterium exchange in isolated proteins relative to the complex provide a topological map of amino acids forming the PPI interface as they are isolated from the deuterated solvent. This approach has been used to determine the binding sites of titin's N2A isoform to the ankyrin repeat protein (Zhou et al., Reference Zhou, Fleming, Lange, Hessel, Bogomolovas, Stronczek, Grundei, Ghassemian, Biju, Borgeson, Bullard, Linke, Chen, Kovermann and Mayans2021a). In that study, the authors showed that the N2A segment, together with its C-terminal IDR linker and the Ig-like domain connected by this linker, constitute the binding site for the ankyrin repeat.

CD is another frequently used technique to monitor the association of two species if the isolated and bound states for the proteins exhibit significant changes in secondary structure content. Such applications have been used to probe structural changes following the binding of tropomyosin and tropomodulin (Greenfield et al., Reference Greenfield, Kostyukova and Hitchcock-DeGregori2005; Kostyukova et al., Reference Kostyukova, Hitchcock-DeGregori and Greenfield2007; Uversky et al., Reference Uversky, Shah, Gritsyna, Hitchcock-DeGregori and Kostyukova2011).

Computational approaches

Bioinformatics approaches

Beyond IDRs ensemble properties, their functions, and especially the binding propensities, are encoded in their sequences (Meng et al., Reference Meng, Uversky and Kurgan2017). Empirical observations indicate that regions within IDRs that are predicted to be more ordered can often fold when bound to targets (Oldfield et al., Reference Oldfield, Cheng, Cortese, Romero, Uversky and Dunker2005). Correspondingly, IDPs are reported to be enriched in molecule recognition elements (MoREs) that serve as target recognition elements and undergo disorder-to-order transitions upon binding (Yang et al., Reference Yang, Gao, Xiong, Su and Huang2019). SLIMs are equally important in mediating PPIs and are often disordered as well (Davey et al., Reference Davey, Van Roey, Weatheritt, Toedt, Uyar, Altenberg, Budd, Diella, Dinkel and Gibson2012). The availability of large, annotated IDP datasets and the rapid development of machine learning techniques have resulted in tens of bioinformatic tools for predicting IDPs' binding sites and propensities for other proteins and nucleic acids (reviewed in Meng et al., Reference Meng, Uversky and Kurgan2017). A recent study demonstrated that MoREs and SLIMs in IDPs are amenable to language-processing techniques and are therefore expected to be more readily detected from IDP amino acid sequences (Lindorff-Larsen and Kragelund, Reference Lindorff-Larsen and Kragelund2021). In addition to reasonably accurate predictive models (Meng et al., Reference Meng, Uversky and Kurgan2017), the rapid and new developments in this area will provide powerful tools to investigate MAPIDs' functions .

Implicit and semi-analytic representations

A number of implicit models have been developed for predicting binding affinities and rates for IDRs. Fundamentally, binding affinities can be defined by the relationship:

(54)$$K_D = \displaystyle{{\,p_D} \over {\,p_A}}$$
(55)$$ = \displaystyle{{\vert \Omega \vert ^{{-}1}\exp ( {-}U_D/k_BT) } \over {\vert \Omega \vert ^{{-}1}\exp ( {-}U_A/k_BT) }}$$
(56)$$\Rightarrow K_D = \exp [ {-( U_D-U_A) /k_{\rm B}T} ] $$

where U D and U A are the free energy of bound and isolated states, respectively. A more rigorous example of note intuitively describes binding affinity in terms of an effective concentration, similar to that introduced in Eq. (10) (Van Valen et al., Reference Van Valen, Haataja and Phillips2009). This model assumes a protein contains two domains linked by an IDR, for which each domain binds to a distinct site on a target (see Fig. 9). If the affinities are K D1 and K D2, the combined affinity can be estimated as:

(57)$$K_D = \displaystyle{{K_{D1}K_{D2}} \over {c_{{\rm eff}}}}$$

where c eff is an effective concentration (Zhou, Reference Zhou2001). For this example, the effective concentration can be determined by an end-to-end probability density, p(d), such that c eff = p(d). We used this concept of effective concentration to characterize TnI/TnC interactions (Siddiqui et al., Reference Siddiqui, Tikunova, Walton, Liu, Meyer, de Tombe, Neilson, Kekenes-Huskey, Salhi, Janssen, Biesiadecki and Davis2016). Although we did not investigate PTMs in that study, in principle one could use this model to investigate the effects of S23/S24 phosphorylation of TnI on TnC's Ca2+ K D, which are known to reduce troponin's apparent Ca2+ affinity (Rao et al., Reference Rao, Cheng, Lindert, Wang, Oxenford, McCulloch, McCammon and Regnier2014).

Fig. 9. (a) Binding of IDP to its partner often goes through a coupled-folding-and-binding process (Sugase et al., Reference Sugase, Dyson and Wright2007), in which both the intramolecular conversion kinetics (section ‘Computational methods for predicting intramolecular dynamics of MAPIDs’) of the IDP and its intermolecular association kinetics (section ‘Computational methods for predicting the MAPID co-assembly’) are important. (b) Theoretical frameworks for describing IDP intermolecular kinetics. The Smoluchowski equation is an approximation for a diffusion-limited association rate (Kim et al., Reference Kim, Meng, Yoo and Chung2018). The Van Valen et al. model (Van Valen et al., Reference Van Valen, Haataja and Phillips2009) (biochemistry on a leash) combines IDP-enhanced effective concentrations and competitive binding to describe IDP/target binding (Eqs. (10) and (56)). The fly-casting model (Shoemakeret al., 2000) explains the kinetic advantage of IDP/target assembly through its fast searching for binding partners. The Brownian dynamics (browndye (Huber and McCammon, Reference Huber and McCammon2010)) and the SEEKR (Votapka et al., Reference Votapka, Jagger, Heyneman and Amaro2017) programs both use the Northrup Allison McCammon algorithm (Northrup et al., Reference Northrup, Allison and McCammon1984) provide simulation-based estimates of association kinetics and thermodynamics quantities.

Avidity is another term often used to describe the fuzzy binding interactions between IDPs and their protein partners. IDPs are well-known to bind to proteins through multi-valent interactions through their many linear sequence motifs (MoRE/SLIM) (Davey et al., Reference Davey, Van Roey, Weatheritt, Toedt, Uyar, Altenberg, Budd, Diella, Dinkel and Gibson2012; Harmon et al., Reference Harmon, Holehouse, Rosen and Pappu2017; Yang et al., Reference Yang, Gao, Xiong, Su and Huang2019). In this regard, there likely exist many binding intermediates with variable stoichiometries. The concept of avidity is proposed to describe the effective binding constant for multivalent interactions between two molecules, and can be quantified by Erlendsson and Teilum (Reference Erlendsson and Teilum2021):

(58)$$K_{av} = \displaystyle{{\sum\nolimits_i {[ RL_i] } } \over {[ R_{\,free}] [ L] }}$$

where [R free] and [L] are the concentrations of free state receptor (folded proteins) and ligand (IDPs), respectively. [RL i] is the concentration of an intermediate complex with i ligands bound. This equation illustrates that the binding constant increases with the number of bound ligands or motifs.

For the kinetics of binding we introduce the mass action relationship:

(59)$${\rm A} + {\rm B}\mathrel{\mathop{\kern0pt\rightleftharpoons}\limits_{{\rm k}_ \hbox{-} }^{{\rm k}_{\rm + \ }}} {\rm A}\cdot {\rm B}$$

where k + represents the association rate constant of species B with target A, while the complex dissociation rate constant is given by k . They are related to the equilibrium constant K D:

(60)$$K_D = \displaystyle{{k_-} \over {k_ + }}$$

In the event that A and B are freely diffusing, spherical, and uniformly reactive, the association rate can be described by the Smoluchowski equation:

(61)$$k_ {+}= 4\pi DR$$

where D is the diffusion constant and R is the distance between the substrates' centers of mass. This relationship arises from the diffusion equation:

(62)$$\displaystyle{{d\rho ( t) } \over {dt}} = {-}\nabla \cdot j( t) $$
(63)$$j = -D\nabla \rho $$

where the latter equation describes the substrate flux across an arbitrary boundary, such as near the PPI interface. An area integral of the substrate flux over the reaction boundary, Γa, yields the association rate

(64)$$k_ {+} = \displaystyle{1 \over {\rho _0}}\smallint _{\Gamma _a}{\vector j}\cdot \hat{n}\; d\Gamma $$

where ρ 0 is the substrate concentration far from the molecular complex and $\hat{n}$ is a surface normal along Γa (Kekenes-Huskey et al., Reference Kekenes-Huskey, Gillette, Hake and McCammon2012). For a spherical binding partner, the form:

(65)$$k_ {+} = 4\pi D\left[{\int_{R_0}^\infty {R^{{-}2}} dR} \right]^{{-}1}$$

is commonly used, where R 0 is the radius of a uniformly reactive, spherical protein target (Shoemaker et al., Reference Shoemaker, Portman and Wolynes2000). Evaluation of the integral yields Eq. (61).

To account for a PMF between the binding proteins, u, the Smoluchowsi equation (Eq. (37)) yields a flux

(66)$$j = D( {\nabla \rho + \beta \rho \nabla u} ) $$

that could be used with Eq. (64). If the reactive center and the PMF are centrosymmetric, Eq. (65) becomes

(67)$$k_ {+} = 4\pi D\left[{\int_{R_0}^\infty {R^{{-}2}} \exp [ \beta u( R) ] dR} \right]^{{-}1\vert }$$

(Shoemaker et al., Reference Shoemaker, Portman and Wolynes2000). These Smoluchowski relationships have been used for several studies of IDP association (Dogan et al., Reference Dogan, Jonasson, Andersson and Jemth2015; Kim et al., Reference Kim, Meng, Yoo and Chung2018; Ozmaian and Makarov, Reference Ozmaian and Makarov2019).

To reflect a tethered substrate such as the N-terminal IDR from TnI, arguments similar to those for Eq. (57) may be assumed. Hence, the effective binding rate, k +,eff, of a species A to B to form the complex C can be described via the scheme

(68)$${\rm A} + {\rm B}\mathrel{\mathop{\kern0pt\rightleftharpoons}\limits_{{\rm k}_ \hbox{-} }^{{\rm k}_{\rm + \ }}} {\rm A}\cdot {\rm B}\mathrel{\mathop{\kern0pt\rightleftharpoons}\limits_{{\rm k}_d}^{{\rm k}_a}} {\rm C}$$

for which the first equilibrium represents the diffusional encounter of the proteins from large distances, while the second equilibrium describes the intrinsic rates for forming C from the ‘encounter’ complex A ⋅ B. The effective binding rate constant can then be evaluated as (Van Valen et al., Reference Van Valen, Haataja and Phillips2009)

(69)$$k_{ + , {\rm eff}} = \displaystyle{{k_{ + , 1}k_a^\ast } \over {k_{-, 1} + k_d^\ast }}$$
(70)$$k_a^{\ast}= k_ap( R) $$

The latter equation represents the impact of the linker region on the basal association rate of the ‘untethered’ binding domain. Here, p(R), can be interpreted as an effective concentration, such as the expression introduced earlier in Eq. (10) (Van Valen et al., Reference Van Valen, Haataja and Phillips2009). k +,1 and k −,1 represent the diffusion-influenced association rate constants from the first equilibrium in Eq. (68), while k a and k d reflect the intrinsic association and dissociation rate constants.

An additional approach characterizes IDR association by a fly-casting mechanism, via which an unfolded domain can accelerate binding to a target (Shoemaker et al., Reference Shoemaker, Portman and Wolynes2000). It relies on computing the flux, j a, for a molecule in unfolded, u, and folded, f, states to yield a combined association rate:

(71)$$k_ {+}= j_u M_u + j_f M_f$$

where M a represents the fraction of species in state a and j a follows from Eq. (66). With this model, the authors demonstrated that binding kinetics can be accelerated by casting unfolded segments toward the target that bind with weak affinity, after which the protein can reel itself toward the target by the simultaneous binding and folding of the segments.

Explicit representations

The kinetics of binding an intrinsically disordered species to a target presents a difficult modeling challenge. This is because of the range of spatio-temporal scales necessary to describe the event. In principle, one can simulate the assembly of two binding species explicitly using unbiased molecular dynamic simulations. Recently, this was demonstrated for the binding of the measles virus nucleoprotein IDP to the X domain of the measles virus phosphoprotein using specialized computing resource Anton for hundreds of μs-length simulations (Robustelli et al., Reference Robustelli, Piana, Shaw and Shaw2020). In this capacity, the association kinetics, given sufficient binding events are observed from the unbiased simulations, can be simply estimated by collecting the number of events per unit time in a fixed volume, such as via k + = 1/t fp where t fp is the first passage time of reaching the bound state (Sun and Kekenes-Huskey, Reference Sun and Kekenes-Huskey2021).

The survival probability provides another means for computing the association rate. This can be done by simulating an ensemble of substrates within a closed volume that can collide with a target (Kim and Yethiraj, Reference Kim and Yethiraj2010). This is given by the reaction:

(72)$$\displaystyle{{dS_R} \over {dt}}( t) = {-}k( t ) C_LS_R( t) $$

where S(t) is the survival probability, e.g. the likelihood that the target remains unbound from 0 to time t. The probability of the target remaining unbound is enumerated from many simulation trajectories as a function of time. The k + at t → ∞ is determined by fitting the expression and extrapolating to large t:

(73)$$\displaystyle{{d\ln ( {S_R} ) } \over {dt}} = {-}k( t ) C_L$$

Hence, the rate can be determined from the survival probability for an element of the disordered domain, such as a SLIM, to remain affixed to a binding site. To our knowledge, such approaches have not been applied to IDRs, or at least not for MAPIDs. This approach generally requires a large number of simulations to obtain convergent results.

Alternatives to explicit, brute force simulations can provide more efficient means for extracting binding kinetics. Biased all-atom MD simulations are a more scalable method to estimating binding kinetics, by favoring trajectories most likely to lead to reaction. As an example, weighted ensemble all-atom simulations were used to obtain a sufficient number of association events between the p53 IDP to its partner, for which the association rates were directly calculated based on a two-state model, e.g. Eq. (59) (Zwier et al., Reference Zwier, Pratt, Adelman, Kaus, Zuckerman and Chong2016).

A related approach places substrates on a reaction boundary (such as the b sphere in BD simulations shown in Fig. 9) and evaluates the flux of substrate toward the reactive center based on explicitly simulating binding trajectories. This allows k + to be determined by (Vijaykumar et al., Reference Vijaykumar, Bolhuis and Ten Wolde2016):

(74)$$k_ {+} = \displaystyle{{k_{a, 0}p_{eq}( R) k_D} \over {k_{a, 0}p_{eq}( R) + k_D}}$$
(75)$$ = [ {1-S_{rad}( {t\to \infty \vert R} ) } ] k_D$$

Here k D is used to represent the diffusion-limited association rate between binding partners, k a,0 is an intrinsic association rate, and p eq is the equilibrium distribution of the substrate. They also show that this association rate can be determined from S rad, which is the survival probability that reflects the likelihood that two partners in contact (radially) diffuse toward the bulk instead of binding. Hence, Eq. (75) describes the process of two partners encountering one another at the diffusion-limited association rate, then binding, before dissociation can occur.

The authors also show that k can be determined by the flux of substrate away from the binding center (Vijaykumar et al., Reference Vijaykumar, Bolhuis and Ten Wolde2016)

(76)$$k_{-}= k_{d, 0}S_{rad}( {t\to \infty \vert R} ) $$
(77)$$ = \displaystyle{{\phi _{B, 0}} \over {{\cal H}_B}}P( {r_n\vert r_0} ) $$

Here, k d,0 is an intrinsic dissociation rate, ϕ B,0 quantifies the flux of substrate reaching a binding interface, ${\cal H}_B$ is an indicator if the substrate were recently bound (Vijaykumar et al., Reference Vijaykumar, Bolhuis and Ten Wolde2016), and P(r n|r 0) is the probability that the substrate escapes before being rebound. Formulations that use a biasing force to drive reactions are also proposed, which could additionally enable one to predict dissociation rates from MD simulations (Maximova et al., Reference Maximova, Postnikov, Lavrova, Farafonov and Nerukh2021).

Markov state modeling. MSMs also provide a convenient framework to model the kinetics of assembly, when the RC can be partitioned into a series of states. To relate these kinetics to an association or dissociation rate, the transition probability matrix can be used to compute quantities such as the MFPT or survival probability. When s i and s j represent the bound and dissociated states, the MFPT (⟨tj) for releasing a bound ligand can be determined, if s j is enforced to be an absorbing state. This residence time is inversely proportional to the off-rate, e.g. k = 1/tj. SEEKR is one such approach (Votapka et al., Reference Votapka, Jagger, Heyneman and Amaro2017, Reference Votapka, Stokely, Ojha and Amaro2022) that has used this formalism to determine an association rate of 9 × 108 M−1s−1 for Ca2+ binding to TnC (Votapka and Amaro, Reference Votapka and Amaro2015).

Brownian dynamics. BD offers a helpful compromise in modeling IDP/protein association by representing solvent effects through a friction coefficient that acts on the solute (see Eq. (23)). Estimating the association rate from this formalism entails enumerating the number of successful collisions between two species versus those that result in the two species diffusing apart. While many approaches are available for estimating this rate, we refer to the frequently used Northrup Allison McCammon algorithm (Northrup et al., Reference Northrup, Allison and McCammon1984) (see Fig. 9).

(78)$$k_{\rm + \ }{\rm} = k_{{\rm + , 1}} p$$

where b is the sphere radius at which the interactions between the reactants become centrosymmetric, k +,1 is the rate constant of arriving at b, and p is the probability that the reactants bind, instead of dissociating. To calculate p, the term β is defined to represent the probability of two reactants reaching b and having a single collision. This leads to two consequences: (1) bind/react with probability α, or (2) escape. For the escaped reactant, there is a probability Δ to re-collide (encounter the b sphere once again). This process is repeated to obtain p by (Northrup et al., Reference Northrup, Allison and McCammon1984):

(79)$$p = \beta _\infty \alpha + \beta _\infty ( 1-\alpha ) \Delta _\infty \alpha + \beta _\infty ( 1-\alpha ) ^2\Delta _\infty ^2 \alpha + \ldots = \displaystyle{{\beta _\infty \alpha } \over {1-( 1-\alpha ) \Delta _\infty }}$$

where α is the probability for the reactants to bind/react after one collision. Details of this derivation are provided in section S2. This approximation is used in the BD code browndye (Huber and McCammon, Reference Huber and McCammon2010). We leveraged this relationship to perform association calculations between an IDR of CN with CaM to demonstrate that increased electrostatic screening due to ionic species (see Eq. (9)) slows the diffusion rate (Sun et al., Reference Sun, Cook, Creamer and Kekenes-Huskey2018).

BD has also been used to show how an IDR accelerates the binding of EtsΔ 138 protein to ERK2 protein (Misiura and Kolomeisky, Reference Misiura and Kolomeisky2020). Specifically, the EtsΔ138 has two binding sites for ERK2, one in its folded domain, and another in an IDR domain tethered to the folded domain. BD simulation showed that binding the IDR site accelerates the overall association rate. Increasing the IDR length amplifies the acceleration effect up to ~4-fold (Misiura and Kolomeisky, Reference Misiura and Kolomeisky2020).

Multi-scale approaches

AAMD and CG techniques have also been widely applied to study IDP binding processes (Liu et al., Reference Liu, Chu, Lu and Wang2017; Chu et al., Reference Chu, Shammas and Wang2020; Sun and Kekenes-Huskey, Reference Sun and Kekenes-Huskey2021). For instance, in Wang et al. (Reference Wang, Chu, Longhi, Roche, Han, Wang and Wang2013), AAMD was used to reveal the free energy landscape of a 20-residue IDR from measles virus nucleoprotein and its binding mechanism with a protein partner. To accelerate the simulation of the binding process, a structure-based potential was used to augment the conventional all-atom AMBER FF99SB-ILDN potentials to bias the sampling toward the complex state (Wang et al., Reference Wang, Chu, Longhi, Roche, Han, Wang and Wang2013). These atomic simulations show that binding occurs through a coupled-folding-and-binding process that consists of an initial complex formation via a conformational selection mechanism and a subsequent downhill induced fitting step (Wang et al., Reference Wang, Chu, Longhi, Roche, Han, Wang and Wang2013). Using a biasing principle, the structure-based model (SBM) is a CG technique that drives binding to restore interprotein contacts present in the native complex. In Chu et al. (Reference Chu, Shammas and Wang2020), SBM CG simulations were used to reveal the role of non-native electrostatics in an IDP/receptor encounter complex. Liu et al. (Reference Liu, Chu, Lu and Wang2017) used a double-bell-potential SBM to reveal that skeletal muscle myosin light chain kinase and CaM bind through a process utilizing both induced fit and conformational selection mechanisms.

When two binding species are well-separated, details of solute–solvent interactions are less important. Electrostatic interactions, however, remain important and influence different stages of an IDR's binding process. Sun et al. (Reference Sun, Cook, Creamer and Kekenes-Huskey2018) utilized BD for well-separated species and AAMD when proteins were loosely bound. This approach revealed that electrostatics both drive pCaN toward CaM and determine the conformation exchange kinetics of the isolated IDR ensemble. Another example is from Chu et al. (Reference Chu, Wang, Gan, Bai, Han, Wang and Wang2012) in which SBM coupled with Debye–Huckel theory was used to evaluate the role of electrostatics in IDP binding. Their simulations demonstrated that electrostatic interactions initially accelerate binding, after which they hinder transitions to the native complex. In these simulations, as the diffusing ensemble of IDR conformations approaches the target, the binding process entails significant refolding to form the final complex. Estimating the kinetics of refolding likely requires atomistic representations.

In the event that two species fold via a conformation selection mechanism, it is possible to estimate the association rate by combining the ‘loose encounter’ association rate with the kinetics of exposing the (IDR) active site, such as a SLIM. Here it is assumed that the IDR has two states, the active state that is capable of binding the partner, and the inactive state in which the binding site is concealed. The rate constants k f and k b reflect the gating kinetics between these two states. The effective association rate for this situation can be estimated via a gating model developed by Szabo et al. (Reference Szabo, Shoup, Northrup and McCammon1982):

(80)$$k_{{\rm eff}}^ {+} = \displaystyle{{k_ + k_{eq}k_bZ[ k_f + k_b] } \over {k_f( k_{eq} + k_ + Z[ k_f + k_b] ) + k_bZ[ k_f + k_b] ( k_ + { + } k_{eq}) }}$$

with

(81)$$Z[ k_f + k_b] = 1 + ( ( k_f + k_b) R^2/D) ^{1/2}$$

where k + is the association rate when the ligand (IDR) is locked into the active state. k eq is a characteristic constant indicating the extent to which the association is diffusion-controlled (see Szabo et al. (Reference Szabo, Shoup, Northrup and McCammon1982) for more details). In Sun et al. (Reference Sun, Cook, Creamer and Kekenes-Huskey2018), we showed using MSMs that the gating of a phosphatase IDR was rapid, which minimized the effect of conformational gating on the association rate. In this way, dynamic ensembles can achieve association kinetics that rival folded proteins.

Current limitations and future outlook

Limitations

Force field accuracy

The numerous IDR modeling approaches described in our review invite an extensive list of limitations and directions for improvement that would benefit MAPID characterization. However, we focus our discussions on atomistic-resolution simulations, which we believe are primed to capture both global structural properties of IDRs and their local attributes. In this regard, improving the force field accuracy for IDP simulations remains one of the foremost challenges in IDR modeling. The flat PES renders IDPs sensitive to force field inaccuracies, as minor errors can drive sampling far from their native ensembles (Pietrek et al., Reference Pietrek, Stelzl and Hummer2020). Here, one key avenue to improve all-atom force field accuracy for IDP simulations rests with protein–water interactions (solvation) and backbone dihedral terms, as tuning these interactions leads to improved agreement between simulations and experiments (Best et al., Reference Best, Zheng and Mittal2014; Song et al., Reference Song, Luo and Chen2017). Recent efforts aiming to improve models for continuum solvation (i.e. ABSINTH (Vitalis and Pappu, Reference Vitalis and Pappu2009)) or capturing main chain interactions via pseudo-improper-dihedral terms (Mioduszewski et al., Reference Mioduszewski, Rycki and Cieplak2020) for IDPs are also promising. Developing accurate IDP force fields is an ongoing challenge, and we refer readers to excellent reviews covering the use of experimental data in IDP force field development (Chan-Yao-Chong et al., Reference Chan-Yao-Chong, Durand and Ha-Duong2019) and strategies for improving IDP force field accuracy (Mu et al., Reference Mu, Liu, Zhang, Luo and Chen2021).

Since the IDRs discussed in this review are adjacent in sequence to well-folded proteins, a secondary goal in IDP force field development is to preserve force field accuracy for the folded constituents. As it has been increasingly realized that many proteins contain both folded and disordered components, developing force fields that are accurate for both IDPs and folded proteins is necessary. Mainstream all-atom force fields such as those from Amber (Weiner et al., Reference Weiner, Kollman, Case, Singh, Ghio, Alagona, Profeta and Weiner1984) or CHARMM (MacKerell et al., Reference MacKerell, Bashford, Bellott, Dunbrack, Evanseck, Field, Fischer, Gao, Guo, Ha, Joseph-McCarthy, Kuchnir, Kuczera, Lau, Mattos, Michnick, Ngo, Nguyen, Prodhom, Reiher, Roux, Schlenkrich, Smith, Stote, Straub, Watanabe, Wiórkiewicz-Kuczera, Yin and Karplus1998) initially had difficulties in simultaneously characterizing IDP and folded proteins, but refinement of the parameters was shown to be helpful (Robustelli et al., Reference Robustelli, Piana and Shaw2018). Similar developments are in progress for CG force fields like maximum entropy optimized force field (MOFF) (Latham and Zhang, Reference Latham and Zhang2021), which was parameterized against experimentally measured radius of gyration for IDPs and the folded structures of proteins. This was found to work well for both IDP and folded proteins (Latham and Zhang, Reference Latham and Zhang2021).

Simulation approaches

Simulation techniques must evolve in parallel with force field developments. It is increasingly clear that ensemble properties of isolated IDPs can be characterized reasonably well by MD simulations coupled with advanced sampling techniques such as REST, HREMD, or with the help of experimentally guided constraints. However, IDP systems with higher complexity and degrees of freedom, such as IDP-mediated LLPS (Garaizar et al., Reference Garaizar, Sanchez-Burgos, Collepardo-Guevara and Espinosa2020; Shea et al., Reference Shea, Best and Mittal2021), IDP-based extracellular matrix (Clarke and Pappu, Reference Clarke and Pappu2017) formation, and IDP-aggregation (Strodel, Reference Strodel2021) still would benefit from more advanced computational resources. This can take the form of advanced hardware and chip architectures, much like the recent GPU revolution in MD simulation. Algorithmic changes, such as the recent adoption of hydrogen mass reweighting to increase simulation step size (Hopkins et al., Reference Hopkins, Le Grand, Walker and Roitberg2015), should be developed in tandem with hardware advances.

Structural data

In addition to modeling developments that could improve IDR modeling as a whole, there are new frontiers specific to the characterization and prediction of MAPID properties. At the time of this writing, there are few reports describing the application of computational techniques to the IDRs of the MAPIDs in Table 1, which renders difficult the probing of the molecular mechanisms driving myofilament function. A key barrier to these applications has been the lack of structural data for the well-folded portions of many of these proteins in isolation, much less in macromolecular complexes like troponin. The configurations of the globular proteins provide important boundary conditions for disordered domains that are tethered to or between domains. Furthermore, knowledge of the higher-order structure of macromolecular complexes is necessary to define the environment, within which an IDR ensemble samples. Advances in cryo-EM spectroscopy and SAXS studies of filament and Z-disk constructions (Sponga et al., Reference Sponga, Arolas, Schwarz, Jeffries, Chamorro, Kostan, Ghisleni, Drepper, Polyansky, Ribeiro, Pedron, Zawadzka-Kazimierczuk, Mlynek, Peterbauer, Doto, Schreiner, Hollerl, Mateos, Geist, Faulkner, Kozminski, Svergun, Warscheid, Zagrovic, Gautel, Konrat and Djinovi-Carugo2021; Wang et al., Reference Wang, Grange, Wagner, Kho, Gautel and Raunser2021) are quickly closing this gap. Furthermore, recent advances in ab initio protein structure programs such as alphafold and rosettafold are providing reliable structures for well-folded domains (Baek et al., Reference Baek, DiMaio, Anishchenko, Dauparas, Ovchinnikov, Lee, Wang, Cong, Kinch, Schaeffer, Millán, Park, Adams, Glassman, DeGiovanni, Pereira, Rodrigues, Dijk, Ebrecht, Opperman, Sagmeister, Buhlheller, Pavkov-Keller, Rathinaswamy, Dalwadi, Yip, Burke, Garcia, Grishin, Adams, Read and Baker2021; Jumper et al., Reference Jumper, Evans, Pritzel, Green, Figurnov, Ronneberger, Tunyasuvunakool, Bates, Žídek, Potapenko, Bridgland, Meyer, Kohl, Ballard, Cowie, Romera-Paredes, Nikolov, Jain, Adler, Back, Petersen, Reiman, Clancy, Zielinski, Steinegger, Pacholska, Berghammer, Bodenstein, Silver, Vinyals, Senior, Kavukcuoglu, Kohli and Hassabis2021). In time, these ab initio programs are likely to achieve similar gains with IDRs of at least isolated proteins. Strategic combinations of experimental and computational advances may begin to realize three-dimensional atlases of myofilament complexes. Efforts underway in the Tardiff and Schwarz labs to assemble thin filament proteins based on cryo-EM structural data (Mason et al., Reference Mason, Lynn, Baldo, Deranek, Tardiff and Schwartz2021; Deranek et al., Reference Deranek, Baldo, Lynn, Schwartz and Tardiff2022) are already bringing this goal into fruition.

In a similar regard, consideration of auxiliary proteins with IDRs that regulate myofilament function will be essential to understand how the myofilaments adapt to physiological (exercise, pregnancy) and pathological influences (elevated blood pressure, myocardial infarction) (Cornwell et al., Reference Cornwell, Li, Sellak, Miller and Word2001; Hamdani et al., Reference Hamdani, de Waard, Messer, Boontje, Kooij, van Dijk, Versteilen, Lamberts, Merkus, Remedios, Duncker, Borbely, Papp, Paulus, Stienen, Marston and van de Velden2008). As an example, CaMKII and CN are both implicated in pathological cardiac remodeling, predominantly due to impacting gene transcription (Maier and Bers, Reference Maier and Bers2002). However, CaMKII is additionally reported to phosphorylate titin (Hidalgo et al., Reference Hidalgo, Chung, Saripalli, Methawasin, Hutchinson, Tsaprailis, Labeit, Mattiazzi and Granzier2013), MyBPC3, and TnI (Tong et al., Reference Tong, Gaffin, Zawieja and Muthuchamy2004), which likely impacts cardiac contractility on a much more rapid basis (Huke and Bers, Reference Huke and Bers2007). Meanwhile, Calcineurin is tethered to the Z-disk via calsarcin; and given the fundamental importance of the Z-disk to mechanosignal transduction (Russell and Solís, Reference Russell and Solís2021), an intriguing possibility is that the phosphatase may directly impact myofilament function through local PTMs. Both proteins also contain IDRs (Rumi-Masante et al., Reference Rumi-Masante, Rusinga, Lester, Dunlap, Williams, Dunker, Weis and Creamer2012; Bhattacharyya et al., Reference Bhattacharyya, Lee, Muratcioglu, Qiu, Nyayapati, Schulman, Groves and Kuriyan2020) that regulate their enzymatic functions, which invites the application of the IDR modeling techniques we discuss in this review. Additional proteins include protein kinase A, protein kinase C, and S100 proteins that are already well-established to impact myofilament function (de Tombe, Reference de Tombe2003; Duarte-Costa et al., Reference Duarte-Costa, Castro-Ferreira, Neves and Leite-Moreira2014).

Lastly, the computational methods we discuss in this review almost exclusively deal with dilute environments that are an idealized representation of the cell cytoplasm. Determining how these proteins behave in cellular structures like the sarcomere will also be important for understanding myofilament function and kinetics, as well as the availability of substrates such as ATP for force generation.

Open questions

Aside from the challenges we outlined in the previous section, there are a number of open questions with the potential to be resolved via computer simulation. These include what fraction of IDRs in the myofilament serves important functional or regulatory roles? Is there an evolutionary basis for why IDRs are broadly distributed throughout myofilament proteins? The high propensity of IDRs estimated via PONDR would suggest that there are many molecular mechanisms modulating myofilament function, so are they all necessary or merely redundant? In addition, is it possible to determine if only a subset of available conformations in an ensemble is important to function? To what extent do the conformational ensembles of isolated proteins or complexes behave similarly to their counterparts in higher-order complexes such as the troponin complex, the Z-disk, or the myofibril? How do we identify which experiments would yield the most helpful information for characterizing conformation ensembles and constraining conformations for modeling? Among the unique challenges of myofilaments introduced in the section ‘General challenges in characterizing native IDR structures’, is the state-of-the-art in experimental characterization sufficient to determine the important mechanisms of myofilament contraction or are advances still needed? For diseases that have causative origins in single nucleotide polymorphisms located in IDRs, can the severity of an IDR-localized mutation or its potential to be therapeutically corrected be estimated?

Future opportunities

Artificial intelligence-based predictions of MAPID IDRs

Atomistic simulations in particular have made considerable advances in enabling the prediction of IDP ensembles. Nonetheless, given the complexity of these proteins and their rapid dynamics, routine, brute force simulations of large disordered regions will remain out of reach for some time. Here machine learning approaches that have demonstrated success in predicting the structures of globular proteins, such as alphafold (Jumper et al., Reference Jumper, Evans, Pritzel, Green, Figurnov, Ronneberger, Tunyasuvunakool, Bates, Žídek, Potapenko, Bridgland, Meyer, Kohl, Ballard, Cowie, Romera-Paredes, Nikolov, Jain, Adler, Back, Petersen, Reiman, Clancy, Zielinski, Steinegger, Pacholska, Berghammer, Bodenstein, Silver, Vinyals, Senior, Kavukcuoglu, Kohli and Hassabis2021) and rosettafold (Baek et al., Reference Baek, DiMaio, Anishchenko, Dauparas, Ovchinnikov, Lee, Wang, Cong, Kinch, Schaeffer, Millán, Park, Adams, Glassman, DeGiovanni, Pereira, Rodrigues, Dijk, Ebrecht, Opperman, Sagmeister, Buhlheller, Pavkov-Keller, Rathinaswamy, Dalwadi, Yip, Burke, Garcia, Grishin, Adams, Read and Baker2021), will likely follow suit with IDPs. As a recent example (Gupta et al., Reference Gupta, Dey, Hicks and Zhou2022), ‘autoencoders’ were informed from short MD simulations to predict NMR chemical shift and SAXS data. As SAXS and shift data become increasingly available for MAPIDs, such autoencoders could be retrained or refined to improve predictions for myofilament proteins. It also has been proposed that an IDR's sequence encodes more information than just their ensemble properties (Lindorff-Larsen and Kragelund, Reference Lindorff-Larsen and Kragelund2021). Given that more than 2000 SLIMs have been identified in IDRs, their interactions to form complex, condensates, and disease-related variants are also likely to be predicted from the sequences with advanced machine learning techniques (Lindorff-Larsen and Kragelund, Reference Lindorff-Larsen and Kragelund2021).

Multi-modal structure determination and modeling

In the last decade, advances in cryo-EM have afforded high-resolution structures of macromolecular structures (Bai et al., Reference Bai, McMullan and Scheres2015). A prime example is the near complete model of the thin filament (Yamada et al., Reference Yamada, Namba and Fujii2020), as well as the major constituents of the thick filament (Daneshparvar et al., Reference Daneshparvar, Taylor, O’Leary, Rahmani, Abbasiyeganeh, Previs and Taylor2020). These represent remarkable advancements toward the ambitious goal of a complete reconstruction of the entire sarcomere. Budding efforts to image the Z-disk to similar resolutions are ongoing (Wang et al., Reference Wang, Grange, Wagner, Kho, Gautel and Raunser2021) and thus it represents the last significant, and likely most challenging, macromolecular myofilament complex to resolve. From a modeling standpoint, the availability of comprehensive MAPID assemblies will present new challenges. The immense computational expense of MD simulations already renders difficult all-atom simulations of isolated proteins at biologically relevant timescales.

Simulations of the full cardiac thin filament of ~880 kD comprising actin monomers, the troponin complex, and tropomyosin are limited to tens of ns (Mason et al., Reference Mason, Lynn, Baldo, Deranek, Tardiff and Schwartz2021; Deranek et al., Reference Deranek, Baldo, Lynn, Schwartz and Tardiff2022). Hence, these detailed simulations represent a billionth of the duration of a typical 1 s heart beat. For this reason, precise questions and judicious choice of modeling approaches are important considerations when seeking to extrapolate molecular predictions to myofilament function. Especially as more structural data of the myofilament become available, further development of multiscale modeling approaches will be necessary to accommodate predictions from disparate simulation approaches suitable for different system sizes and timescales. Here, efforts to reconstruct dynamic models of whole-heart function from cell-based descriptors (Trayanova and Rice, Reference Trayanova and Rice2011; Niederer et al., Reference Niederer, Campbell and Campbell2019; Timmermann et al., Reference Timmermann, Edwards, Wall, Sundnes and McCulloch2019), and myofilament contraction from filament-level molecular interactions (Fenwick et al., Reference Fenwick, Wood and Tanner2017; Powers et al., Reference Powers, Yuan, McCabe, Murray, Childers, Flint, Moussavi-Harami, Mohran, Castillo, Zuzek, Ma, Daggett, McCulloch, Irving and Regnier2019; de Winter et al., Reference de Winter, Molenaar, Yuen, van der Pijl, Shen, Conijn, van de Locht, Willigenburg, Bogaards, van Kleef, Lassche, Persson, Rassier, Sztal, Ruparelia, Oorschot, Ramm, Hall, Xiong, Johnson, Li, Kiss, Lozano-Vidal, Boon, Marabita, Nogara, Blaauw, Rodenburg, Küsters, Doorduin, Beggs, Granzier, Campbell, Ma, Irving, Malfatti, Romero, Bryson-Richardson, van Engelen, Voermans and Ottenheijm2020; Creso and Campbell, Reference Creso and Campbell2021; Sharifi et al., Reference Sharifi, Mann, Rockward, Mehri, Mojumder, Lee, Campbell and Wenk2021; Kosta et al., Reference Kosta, Colli, Ye and Campbell2022), could provide essential guidance toward the incorporation of molecular-resolution data from a variety of MAPIDs. Importantly, these approaches represent the first steps toward combining different data sets that can be conflicting and overlapping at different resolutions. Generalized approaches such as the stochastic multiscale model comprising the coarse-graining of atomic structures followed by BD and Langevin dynamics (Aboelkassem et al., Reference Aboelkassem, McCabe, Huber, Regnier, McCammon and McCulloch2019), could lessen the user bias in how structures, models, and modeling results are fused. Similarly, Bayesian approaches could be used to help determine model uncertainties for multi-resolution data and where to prioritize data collection (Liu et al., Reference Liu, Shi, Daumé and Voth2008).

Liquid–liquid phase transition (LLPS) mediated MAPID functions

Biomolecules can condense into liquid phases, forming membrane-less organelles or droplets, through the LLPS process (Shea et al., Reference Shea, Best and Mittal2021). LLPS domains are increasingly of interest as they provide unique microenvironments for biological processes, such as forming the replication machinery for viruses (Savastano et al., Reference Savastano, Ibáñez, Rankovic and Zweckstetter2020), providing platforms for protein interactions (Nott et al., Reference Nott, Craggs and Baldwin2016), and enriching substrate concentrations (O'Flynn and Mittag, Reference O'Flynn and Mittag2021). IDPs are major players in LLPS, because IDPs tend to form multi-valent PPIs with themselves or with other proteins, thus driving LLPS (Harmon et al., Reference Harmon, Holehouse, Rosen and Pappu2017). In the sarcomere, LLPS may also serve an important role in mediating protein interactions. For instance, Sponga et al. showed that in the sarcomere, the IDP protein FATZ-1 condenses into a liquid phase that may provide a mechanism for its interaction with α-actinin (Sponga et al., Reference Sponga, Arolas, Schwarz, Jeffries, Chamorro, Kostan, Ghisleni, Drepper, Polyansky, Ribeiro, Pedron, Zawadzka-Kazimierczuk, Mlynek, Peterbauer, Doto, Schreiner, Hollerl, Mateos, Geist, Faulkner, Kozminski, Svergun, Warscheid, Zagrovic, Gautel, Konrat and Djinovi-Carugo2021). Understanding LLPS in the context of the myofilament is important but is in its infancy, in part because characterizations of MAPIDs are incomplete, and the simulation techniques needed to describe this condensed matter phenomenon are non-trivial. This will improve with more experimental characterizations of MAPIDs, and computational tool development.

Genetics

Ultimately, a compelling motivation for understanding the detailed molecular mechanisms of myofilament function is to contextualize missense variants of myofilament genes and tailor therapeutic intervention. Prominent gene databases such as ClinVar (Landrum et al., Reference Landrum, Lee, Riley, Jang, Rubinstein, Church and Maglott2013) and GnomeAD (Karczewski et al., Reference Karczewski, Francioli, Tiao, Cummings, Alföldi, Wang, Collins, Laricchia, Ganna, Birnbaum, Gauthier, Brand, Solomonson, Watts, Rhodes, Singer-Berk, England, Seaby, Kosmicki, Walters, Tashman, Farjoun, Banks, Poterba, Wang, Seed, Whiffin, Chong, Samocha, Pierce-Hoffman, Zappala, O'Donnell-Luria, Minikel, Weisburd, Lek, Ware, Vittal, Armean, Bergelson, Cibulskis, Connolly, Covarrubias, Donnelly, Ferriera, Gabriel, Gentry, Gupta, Jeandet, Kaplan, Llanwarne, Munshi, Novod, Petrillo, Roazen, Ruano-Rubio, Saltzman, Schleicher, Soto, Tibbetts, Tolonen, Wade, Talkowski, Neale, Daly and MacArthur2020) have an ever-increasing number of VUSs. The impacts of variants on myofilament function are diverse in phenotype and severity, including the Ca2+ sensitivity of force generation or relaxation (Chung et al., Reference Chung, Biesiadecki, Ziolo, Davis and Janssen2016; Shafaattalab et al., Reference Shafaattalab, Li, Lin, Stevens, Dewar, Lynn, Sanatani, Laksman, Morin, Petegem, Hove-Madsen, Tieleman, Davis and Tibbits2019) that in many cases contribute to cardiac dysfunction (Yar et al., Reference Yar, Monasky and Solaro2014). Efforts to extrapolate these variants to dysfunctional phenotypes via molecular simulations are gaining traction (Shafaattalab et al., Reference Shafaattalab, Li, Lin, Stevens, Dewar, Lynn, Sanatani, Laksman, Morin, Petegem, Hove-Madsen, Tieleman, Davis and Tibbits2019; Mason et al., Reference Mason, Lynn, Baldo, Deranek, Tardiff and Schwartz2021; Sewanan et al., Reference Sewanan, Park, Rynkiewicz, Racca, Papoutsidakis, Schwan, Jacoby, Moore, Lehman, Qyang and Campbell2021).However, detailed simulations of myofilament proteins and especially their IDRs are computationally intensive. Here, machine learning approaches again may help extrapolate experimental and simulation data from known variants to de novo variants (reviewed in Kekenes-Huskey et al., Reference Kekenes-Huskey, Burgess, Sun, Bartos, Rozmus, anderson, January, Eckhardt and Delisle2022). Combining simulations and experimental characterization of known variants could be used, together with bioinformatics approaches, to identify potential druggable protein/protein or protein/SLIM interactions.

With such developments, we may begin to realize the therapeutic potential for controlling or restoring the intrinsic properties of disordered regions containing pathogenic variants. IDRs can have both pathological and helpful properties, therefore, efforts need to target the conformational ensemble members that primarily promote dysfunction (Uversky, Reference Uversky and Renaud2020). Progress has been made in IDR-targeted drug design, as was shown for P53 to Mdm2 (Uversky, Reference Uversky and Renaud2020), a cell cycle regulator, and p27 (Iconaru et al., Reference Iconaru, Ban, Bharatham, Ramanathan, Zhang, Shelat, Zuo and Kriwacki2015), a small molecule that binds to a transient IDR site, and the protein tyrosine phosphatase PTP1B, a small molecule that cooperatively binds an IDR (Krishnan et al., Reference Krishnan, Koveal, Miller, Xue, Akshinthala, Kragelj, Jensen, Gauss, Page, Blackledge, Muthuswamy, Peti and Tonks2014). Many other IDP-targeting small molecules have been reported, such as for the c-MyC transcription factor (reviewed in Ruan et al., Reference Ruan, Sun, Zhang, Liu and Lai2019). These small molecules can achieve binding specificity to an IDP through transient interactions, modulating an IDP's ensemble properties, and altering protein function (Chen et al., Reference Chen, Liu and Chen2020). Thus far, drug design for IDPs has largely targeted specific sites, such as binding pockets or PPIs (Uversky, Reference Uversky and Renaud2020)

Regulatory control

As remarked in section ‘Properties of MAPIDs and their characterization’, phosphorylation represents one of the most frequently studied PTMs in the myofilament proteins. The PKA, PKC, and CaMKII kinases are among the most common kinase targets activated in the myofilament, in response to β-adrenergic, muscarinic, and calcium signaling (Bers, Reference Bers2001). In the myofilament, PKA and PKC primarily target troponin (van der Velden and Stienen, Reference van der Velden and Stienen2019). Interestingly, these kinases also contain IDPs (Akimoto et al., Reference Akimoto, Selvaratnam, McNicholl, Verma, Taylor and Melacini2013; Yang and Igumenova, Reference Yang and Igumenova2013). Opposing the activity of kinases are phosphatases. Phosphatases including protein phosphatase 1, protein phosphatase 2A (van der Velden and Stienen, Reference van der Velden and Stienen2019), and CN (Rumi-Masante et al., Reference Rumi-Masante, Rusinga, Lester, Dunlap, Williams, Dunker, Weis and Creamer2012) frequently regulate myofilament proteins. Although phosphatases commonly assume a folded catalytic domain, their regulatory domains as well as regulator proteins such as spinophilin contain IDRs (Marsh et al., Reference Marsh, Dancheck, Ragusa, Allaire, Forman-Kay and Peti2010; Rumi-Masante et al., Reference Rumi-Masante, Rusinga, Lester, Dunlap, Williams, Dunker, Weis and Creamer2012).

Understanding regulatory mechanisms controlling the myofilament proteins may therefore benefit from analogous studies of IDRs in the presence of these phosphatases and kinases. Conversely, there may be value in recognizing the role of myofilament proteins in modulating regulatory mechanisms. As an example, CN is modulated by the myofilaments through both α-actinin and calsarcin, which compete for binding CN (Frey et al., Reference Frey, Richardson and Olson2000; Seto et al., Reference Seto, Quinlan, Lek, Zheng, Garton, MacArthur, Hogarth, Houweling, Gregorevic, Turner, Cooney, Yang and North2013).

Concluding remarks

The recent decade has unveiled exciting developments in computational and experimental techniques toward resolving the structure and molecular mechanisms of myofilament-associated proteins and their functions. Despite this, only a small fraction of proteins from the myofilament have been resolved at atomistic resolutions. The remaining myofilament proteins have limited structural information. The reviewed topics of IDR structure prediction, ensemble kinetics, and protein co-assembly will undoubtedly provide a basis for characterizing the remaining proteins. However, continued progress toward advancing techniques to overcome many limitations will be essential to mapping gene sequence to function. These advances could help tackle prominent open questions relating to intrinsically disordered proteins that influence myofilament function and dysfunction.

Supplementary material

The supplementary material for this article can be found at https://doi.org/10.1017/S003358352300001X.

Acknowledgements

Our review primarily focused on simulations of myofilament proteins with intrinsically disordered regions within the last several years. We apologize if we have missed germane publications on this topic. PMK-H would like to thank his colleagues for careful review of this text. Research reported in this publication was supported by the Maximizing Investigators' Research Award (MIRA) (R35) from the National Institute of General Medical Sciences (NIGMS) of the National Institutes of Health (NIH) under grant number R35GM124977. This work used the Extreme Science and Engineering Discovery Environment (XSEDE) (Towns et al., Reference Towns, Cockerill, Dahan, Foster, Gaither, Grimshaw, Hazlewood, Lathrop, Lifka, Peterson, Roskies, Scott and Wilkins-Diehr2014), which is supported by the National Science Foundation grant number ACI-1548562.

Footnotes

1 These are ‘protein changing’ variants, downloaded from ClinVar on 6/2/2022. The ‘disease-associated’ variants are those categorized as ‘pathogenic’ with selection criteria provided; variant of unknown significance (VUS) refer to variants that have not yet been characterized

2 Isoforms with spectra count (SC) >10 were selected. The spectra count is defined as the total number of spectra identified for a protein via mass spectrometry.

References

Abdi, KM, Mohler, PJ, Davis, JQ and Bennett, V (2006) Isoform specificity of ankyrin-B: a site in the divergent C-terminal domain is required for intramolecular association. Journal of Biological Chemistry 281, 57415749.CrossRefGoogle ScholarPubMed
Aboelkassem, Y, McCabe, KJ, Huber, GA, Regnier, M, McCammon, JA and McCulloch, AD (2019) A stochastic multiscale model of cardiac thin filament activation using Brownian-Langevin dynamics. Biophysical Journal 117, 22552272.CrossRefGoogle ScholarPubMed
Abyzov, A, Salvi, N, Schneider, R, Maurin, D, Ruigrok, RW, Jensen, MR and Blackledge, M (2016) Identification of dynamic modes in an intrinsically disordered protein using temperature-dependent NMR relaxation. Journal of the American Chemical Society 138, 62406251.CrossRefGoogle Scholar
Agarwal, R, Paulo, JA, Toepfer, CN, Ewoldt, JK, Sundaram, S, Chopra, A, Zhang, Q, Gorham, J, DePalma, SR, Chen, CS, Gygi, SP, Seidman, CE and Seidman, JG (2021) Filamin C cardiomyopathy variants cause protein and lysosome accumulation. Circulation Research 129, 751766.CrossRefGoogle ScholarPubMed
Ahn, S-H, Huber, GA and McCammon, JA (2022) Investigating intrinsically disordered proteins with Brownian dynamics. Frontiers in Molecular Biosciences 9, 898838.CrossRefGoogle ScholarPubMed
Akimoto, M, Selvaratnam, R, McNicholl, ET, Verma, G, Taylor, SS and Melacini, G (2013) Signaling through dynamic linkers as revealed by PKA. Proceedings of the National Academy of Sciences 110, 1423114236.CrossRefGoogle ScholarPubMed
Alamo, L, Wriggers, W, Pinto, A, Bártoli, F, Salazar, L, Zhao, F-Q, Craig, R and Padrón, R (2008) Three-dimensional reconstruction of tarantula myosin filaments suggests how phosphorylation may regulate myosin activity. Journal of Molecular Biology 384, 780797.CrossRefGoogle ScholarPubMed
Alamo, L, Qi, D, Wriggers, W, Pinto, A, Zhu, J, Bilbao, A, Gillilan, RE, Hu, S and Padron, R (2016) Conserved intramolecular interactions maintain myosin interacting-heads motifs explaining tarantula muscle super-relaxed state structural basis. Journal of Molecular Biology 428, 11421164.CrossRefGoogle ScholarPubMed
Alamo, L, Ware, JS, Pinto, A, Gillilan, RE, Seidman, JG, Seidman, CE and Padron, R (2017) Effects of myosin variants on interacting-heads motif explain distinct hypertrophic and dilated cardiomyopathy phenotypes. eLife 6, e24634.CrossRefGoogle ScholarPubMed
Anbo, H, Sato, M, Okoshi, A and Fukuchi, S (2019) Functional segments on intrinsically disordered regions in disease-related proteins. Biomolecules 9, biom9030088.CrossRefGoogle ScholarPubMed
An, X, Debnath, G, Guo, X, Liu, S, Lux, SE, Baines, A, Gratzer, W and Mohandas, N (2005) Identification and functional characterization of protein 4.1R and actin-binding sites in erythrocyte beta spectrin: regulation of the interactions by phosphatidylinositol-4,5-bisphosphate. Biochemistry 44, 1068110688.CrossRefGoogle ScholarPubMed
Appadurai, R, Nagesh, J and Srivastava, A (2021) High resolution ensemble description of metamorphic and intrinsically disordered proteins using an efficient hybrid parallel tempering scheme. Nature Communications 12, 958.CrossRefGoogle ScholarPubMed
Arndt, C, Koristka, S, Bartsch, H and Bachmann, M (2012) Native polyacrylamide gels. Methods in Molecular Biology 869, 4953.CrossRefGoogle ScholarPubMed
Arsiccio, A, Pisano, R and Shea, J-E (2022) A new transfer free energy based implicit solvation model for the description of disordered and folded proteins. The Journal of Physical Chemistry B 126, 61806190.CrossRefGoogle ScholarPubMed
Baek, M, DiMaio, F, Anishchenko, I, Dauparas, J, Ovchinnikov, S, Lee, GR, Wang, J, Cong, Q, Kinch, LN, Schaeffer, RD, Millán, C, Park, H, Adams, C, Glassman, CR, DeGiovanni, A, Pereira, JH, Rodrigues, AV, Dijk, AAvan, Ebrecht, AC, Opperman, DJ, Sagmeister, T, Buhlheller, C, Pavkov-Keller, T, Rathinaswamy, MK, Dalwadi, U, Yip, CK, Burke, JE, Garcia, KC, Grishin, NV, Adams, PD, Read, RJ and Baker, D (2021) Accurate prediction of protein structures and interactions using a three-track neural network. Science 373, 871876.CrossRefGoogle ScholarPubMed
Bai, X-C, McMullan, G and Scheres, SH (2015) How cryo-EM is revolutionizing structural biology. Trends in Biochemical Sciences 40, 4957.CrossRefGoogle ScholarPubMed
Baldwin, AJ, Walsh, P, Hansen, DF, Hilton, GR, Benesch, JLP, Sharpe, S and Kay, LE (2012) Probing dynamic conformations of the high-molecular-weight aB-crystallin heat shock protein ensemble by NMR spectroscopy. Journal of the American Chemical Society 134, 1534315350.CrossRefGoogle Scholar
Ban, D (2020) Evolving role of conformational dynamics in understanding fundamental biomolecular behavior. In Miller, JM (ed.), Mechanistic Enzymology: Bridging Structure and Function. Washington, D.C.: ACS Publications, pp. 5781.CrossRefGoogle Scholar
Bang, M-L and Chen, J (2015) Roles of nebulin family members in the heart. Circulation Journal 79, 20812087.CrossRefGoogle ScholarPubMed
Bang, M-L, Mudry, RE, McElhinny, AS, Trombitas, K, Geach, AJ, Yamasaki, R, Sorimachi, H, Granzier, H, Gregorio, CC and Labeit, S (2001) Myopalladin, a novel 145-kilodalton sarcomeric protein with multiple roles in Z-disc and I-band protein assemblies. Journal of Cell Biology 153, 413428.CrossRefGoogle ScholarPubMed
Banks, A, Qin, S, Weiss, KL, Stanley, CB and Zhou, HX (2018) Intrinsically disordered protein exhibits both compaction and expansion under macromolecular crowding. Biophysical Journal 114, 10671079.CrossRefGoogle ScholarPubMed
Ban, D, Smith, CA, Groot, BLde, Griesinger, C and Lee, D (2017) Recent advances in measuring the kinetics of biomolecules by NMR relaxation dispersion spectroscopy. Archives of Biochemistry and Biophysics 628, 8191.CrossRefGoogle ScholarPubMed
Belus, A, Piroddi, N, Scellini, B, Tesi, C, Amati, GD, Girolami, F, Yacoub, M, Cecchi, F, Olivotto, I and Poggesi, C (2008) The familial hypertrophic cardiomyopathy-associated myosin mutation R403Q accelerates tension generation and relaxation of human cardiac myofibrils. The Journal of Physiology 586, 36393644.CrossRefGoogle ScholarPubMed
Benayad, Z, Blow, SV, Stelzl, LS and Hummer, G (2021) Simulation of FUS protein condensates with an adapted coarse-grained model. Journal of Chemical Theory and Computation 17, 525537.CrossRefGoogle ScholarPubMed
Berezhkovskii, A and Szabo, A (2005) One-dimensional reaction coordinates for diffusive activated rate processes in many dimensions. The Journal of Chemical Physics 122, 014503.CrossRefGoogle ScholarPubMed
Bernadó, P, Blanchard, L, Timmins, P, Marion, D Ruigrok, RWH (2005) A structural model for unfolded proteins from residual dipolar couplings and small-angle x-ray scattering. Proceedings of the National Academy of Sciences 102, 1700217007.CrossRefGoogle ScholarPubMed
Bernadó, P, Mylonas, E, Petoukhov, MV, Blackledge, M and Svergun, DI (2007) Structural characterization of flexible proteins using small-angle X-ray scattering. Journal of the American Chemical Society 129, 56565664.CrossRefGoogle ScholarPubMed
Bernetti, M, Masetti, M, Pietrucci, F, Blackledge, M, Jensen, MR, Recanatini, M, Mollica, L and Cavalli, A (2017) Structural and kinetic characterization of the intrinsically disordered protein SeV NTAIL through enhanced sampling simulations. Journal of Physical Chemistry B 121, 95729582.CrossRefGoogle ScholarPubMed
Bers, DM (2001) Excitation-Contraction Coupling and Cardiac Contractile Force. Ed. by D. M. Bers, Vol. 1. Springer, Dordrecht: Kluwer Academic Publishers, p. 427.CrossRefGoogle Scholar
Best, RB (2017) Computational and theoretical advances in studies of intrinsically disordered proteins. Current Opinion in Structural Biology 42, 147154.CrossRefGoogle ScholarPubMed
Best, RB, Zheng, W and Mittal, J (2014) Balanced protein-water interactions improve properties of disordered proteins and non-specific protein association. Journal of Chemical Theory and Computation 10, 51135124.CrossRefGoogle ScholarPubMed
Bhattacharya, S and Lin, X (2019) Recent advances in computational protocols addressing intrinsically disordered proteins. Biomolecules 9, 146.CrossRefGoogle ScholarPubMed
Bhattacharyya, M, Lee, YK, Muratcioglu, S, Qiu, B, Nyayapati, P, Schulman, H, Groves, JT and Kuriyan, J (2020) Flexible linkers in CaMKII control the balance between activating and inhibitory autophosphorylation. eLife 9, e53670.CrossRefGoogle ScholarPubMed
Biesiadecki, BJ and Westfall, MV (2019) Troponin I modulation of cardiac performance: plasticity in the survival switch. Archives of Biochemistry and Biophysics 664, 914.CrossRefGoogle ScholarPubMed
Boateng, SY, Belin, RJ, Geenen, DL, Margulies, KB, Martin, JL, Hoshijima, M, de Tombe, PP and Russell, B (2007) Cardiac dysfunction and heart failure are associated with abnormalities in the subcellular distribution and amounts of oligomeric muscle LIM protein. American Journal of Physiology-Heart and Circulatory Physiology 292, H259H269.CrossRefGoogle ScholarPubMed
Boczkowska, M, Rebowski, G, Kremneva, E, Lappalainen, P and Dominguez, R (2015) How leiomodin and tropomodulin use a common fold for different actin assembly functions. Nature Communications 6, 8314.CrossRefGoogle ScholarPubMed
Bowerman, S, Curtis, JE, Clayton, J, Brookes, EH and Wereszczynski, J (2019) BEES: Bayesian ensemble estimation from SAS. Biophysical Journal 117, 399407.CrossRefGoogle ScholarPubMed
Bowman, JD and Lindert, S (2019) Computational studies of cardiac and skeletal troponin. Frontiers in Molecular Biosciences 6, 68.CrossRefGoogle ScholarPubMed
Braun, N, Zacharias, M, Peschek, J, Kastenmuller, A, Zou, J, Hanzlik, M, Haslbeck, M, Rappsilber, J, Buchner, J and Weinkauf, S (2011) Multiple molecular architectures of the eye lens chaperone αB-crystallin elucidated by a triple hybrid approach. Proceedings of the National Academy of Sciences 108, 2049120496.CrossRefGoogle ScholarPubMed
Brodehl, A, Dieding, M, Klauke, B, Dec, E, Madaan, S, Huang, T, Gargus, J, Fatima, A, Saric, T, Cakar, H, Walhorn, V, Tonsing, K, Skrzipczyk, T, Cebulla, R, Gerdes, D, Schulz, U, Gummert, J, Svendsen, JH, Olesen, MS, Anselmetti, D, Christensen, AH, Kimonis, V and Milting, H (2013) The novel desmin mutant p.A120D impairs filament formation, prevents intercalated disk localization, and causes sudden cardiac death. Circulation: Cardiovascular Genetics 6, 615623.Google ScholarPubMed
Brutscher, B, Felli, IC, Gil-Caballero, S, Hošek, T, Kümmerle, R, Piai, A, Pierattelli, R and Sólyom, Z (2015) NMR methods for the study of instrinsically disordered proteins structure, dynamics, and interactions: general overview and practical guidelines. In Felli, IC and Pierattelli, R (eds), Intrinsically Disordered Proteins Studied by NMR Spectroscopy. Cham: Springer International Publishing, pp. 49122.CrossRefGoogle Scholar
Carniel, E, Taylor, MR, Sinagra, G, Lenarda, AD, Ku, L, Fain, PR, Boucek, MM, Cavanaugh, J, Miocic, S, Slavov, D, Graw, SL, Feiger, J, Zhu, XZ, Dao, D, Ferguson, DA, Bristow, MR and Mestroni, L (2005) α-Myosin heavy chain. Circulation 112, 5459.CrossRefGoogle ScholarPubMed
Chalovich, JM and Schroeter, MM (2010) Synaptopodin family of natively unfolded, actin binding proteins: physical properties and potential biological functions. Biophysical Reviews 2, 181189.CrossRefGoogle ScholarPubMed
Chan-Yao-Chong, M, Durand, D and Ha-Duong, T (2019) Molecular dynamics simulations combined with nuclear magnetic resonance and/or small-angle X-ray scattering data for characterizing intrinsically disordered protein conformational ensembles. Journal of Chemical Information and Modeling 59, 17431758.CrossRefGoogle ScholarPubMed
Chang, M, Wilson, CJ, Karunatilleke, NC, Moselhy, MH, Karttunen, M and Choy, WY (2021) Exploring the conformational landscape of the Neh4 and Neh5 domains of Nrf2 using two different force fields and circular dichroism. Journal of Chemical Theory and Computation 17, 31453156.CrossRefGoogle ScholarPubMed
Chen, M (2021) Collective variable-based enhanced sampling and machine learning. The European Physical Journal B 94, 211.CrossRefGoogle ScholarPubMed
Chen, J, Liu, X and Chen, J (2020) Targeting intrinsically disordered proteins through dynamic interactions. Biomolecules 10, 743.CrossRefGoogle ScholarPubMed
Cheng, H, Kimura, K, Peter, AK, Cui, L, Ouyang, K, Shen, T, Liu, Y, Gu, Y, Dalton, ND, Evans, SM, Knowlton, KU, Peterson, KL and Chen, J (2010) Loss of enigma homolog protein results in dilated cardiomyopathy. Circulation Research 107, 348356.CrossRefGoogle ScholarPubMed
Cheng, Y, Lindert, S, Kekenes-Huskey, P, Rao, VS, Solaro, RJ, Rosevear, PR, Amaro, R, McCulloch, AD, McCammon, JA and Regnier, M (2014) Computational studies of the effect of the S23D/S24D troponin I mutation on cardiac troponin structural dynamics. Biophysical Journal 107, 16751685.CrossRefGoogle ScholarPubMed
Cheng, Y, Rao, V, Tu, A-y, Lindert, S, Wang, D, Oxenford, L, McCulloch, AD, McCammon, JA and Regnier, M (2015) Troponin I mutations R146G and R21C alter cardiac troponin function, contractile properties, and modulation by protein kinase A (PKA)-mediated phosphorylation*. Journal of Biological Chemistry 290, 2774927766.CrossRefGoogle ScholarPubMed
Chen, K, Li, J, Wang, C, Wei, Z, and Zhang, M (2017) Autoinhibition of ankyrin-B/G membrane target bindings by intrinsically disordered segments from the tail regions. eLife 6, e29150.CrossRefGoogle ScholarPubMed
Cheung, MS, García, AE and Onuchic, JN (2002) Protein folding mediated by solvation: water expulsion and formation of the hydrophobic core occur after the structural collapse. Proceedings of the National Academy of Sciences 99, 685690.CrossRefGoogle ScholarPubMed
Chiappori, F, Mattiazzi, L, Milanesi, L and Merelli, I (2016) A novel molecular dynamics approach to evaluate the effect of phosphorylation on multimeric protein interface: the αB-crystallin case study. BMC Bioinformatics 17, 57.CrossRefGoogle ScholarPubMed
Chopra, A, Kutys, ML, Zhang, K, Polacheck, WJ, Sheng, CC, Luu, RJ, Eyckmans, J, Hinson, JT, Seidman, JG, Seidman, CE and Chen, CS (2018) Force generation via B-cardiac myosin, titin, and a-actinin drives cardiac sarcomere assembly from cell-matrix adhesions. Developmental Cell 44, 8796.e5.CrossRefGoogle Scholar
Cho, HS, Schotte, F, Stadnytskyi, V and Anfinrud, P (2021) Time-resolved X-ray scattering studies of proteins. Current Opinion in Structural Biology 70, 99107.CrossRefGoogle ScholarPubMed
Chu, X and Wang, J (2019) Position-, disorder-, and salt-dependent diffusion in binding-coupled-folding of intrinsically disordered proteins. Physical Chemistry Chemical Physics 21, 56345645.CrossRefGoogle ScholarPubMed
Chu, WT, Shammas, SL and Wang, J (2020) Charge interactions modulate the encounter complex ensemble of two differently charged disordered protein partners of KIX. Journal of Chemical Theory and Computation 16, 38563868.CrossRefGoogle ScholarPubMed
Chung, J-H, Biesiadecki, BJ, Ziolo, MT, Davis, JP and Janssen, PML (2016) Myofilament calcium sensitivity: role in regulation of in vivo cardiac contraction and relaxation. Frontiers in Physiology 7, 562.CrossRefGoogle ScholarPubMed
Chu, P-H, Bardwell, WM, Gu, Y, Ross, J and Chen, J (2000) FHL2 (SLIM3) is not essential for cardiac development and function. Molecular and Cellular Biology 20, 74607462.CrossRefGoogle Scholar
Chu, X, Wang, Y, Gan, L, Bai, Y, Han, W, Wang, E and Wang, J (2012) Importance of electrostatic interactions in the association of intrinsically disordered histone chaperone Chz1 and histone H2A.Z-H2B. PLoS Computational Biology 8, e1002608.CrossRefGoogle ScholarPubMed
Chvez-Garca, C, Hnin, J and Karttunen, M (2022) Multiscale computational study of the conformation of the full-length intrinsically disordered protein MeCP2. Journal of Chemical Information and Modeling 62, acs.jcim.1c01354.Google Scholar
Clapham, DE (2007) Calcium signaling. Cell 131, 10471058.CrossRefGoogle ScholarPubMed
Clarke, J and Pappu, RV (2017) Editorial overview: protein folding and binding, complexity comes of age. Current Opinion in Structural Biology 42, vvii.CrossRefGoogle ScholarPubMed
Colpan, M, Ly, T, Grover, S, Tolkatchev, D and Kostyukova, AS (2017) The cardiomyopathy-associated K15N mutation in tropomyosin alters actin filament pointed end dynamics. Archives of Biochemistry and Biophysics 630, 1826.CrossRefGoogle ScholarPubMed
Colson, BA, Thompson, AR, Espinoza-Fonseca, LM and Thomas, DD (2016) Site-directed spectroscopy of cardiac myosin-binding protein C reveals effects of phosphorylation on protein structural dynamics. Proceedings of the National Academy of Sciences 113, 32333238.CrossRefGoogle ScholarPubMed
Cool, AM and Lindert, S (2021) Computational methods elucidate consequences of mutations and post-translational modifications on troponin I effective concentration to troponin C. The Journal of Physical Chemistry B 125, 73887396.CrossRefGoogle ScholarPubMed
Cornwell, TL, Li, J, Sellak, H, Miller, RT and Word, RA (2001) Reorganization of myofilament proteins and decreased cGMP-dependent protein kinase in the human uterus during pregnancy. The Journal of Clinical Endocrinology & Metabolism 86, 39813988.CrossRefGoogle ScholarPubMed
Creso, JG and Campbell, SG (2021) Potential impacts of the cardiac troponin I mobile domain on myofilament activation and relaxation. Journal of Molecular and Cellular Cardiology 155, 5057.CrossRefGoogle ScholarPubMed
Crocini, C and Gotthardt, M (2021) Cardiac sarcomere mechanics in health and disease. Biophysical Reviews 13, 637652.CrossRefGoogle ScholarPubMed
Cunha, SR and Mohler, PJ (2008) Obscurin targets ankyrin-B and protein phosphatase 2A to the cardiac M-line*. Journal of Biological Chemistry 283, 3196831980.CrossRefGoogle Scholar
Dai, W, Sengupta, AM and Levy, RM (2015) First passage times, lifetimes, and relaxation times of unfolded proteins. Physical Review Letters 115, 048101.CrossRefGoogle ScholarPubMed
Daneshparvar, N, Taylor, DW, O’Leary, TS, Rahmani, H, Abbasiyeganeh, F, Previs, MJ and Taylor, KA (2020) CryoEM structure of Drosophila flight muscle thick filaments at 7 Å resolution. Life Science Alliance 3, e202000823.CrossRefGoogle ScholarPubMed
Das, RK and Pappu, RV (2013) Conformations of intrinsically disordered proteins are influenced by linear sequence distributions of oppositely charged residues. Proceedings of the National Academy of Sciences 110, 1339213397.CrossRefGoogle ScholarPubMed
Das, RK, Ruff, KM and Pappu, RV (2015) Relating sequence encoded information to form and function of intrinsically disordered proteins. Current Opinion in Structural Biology 32, 102112.CrossRefGoogle ScholarPubMed
Das, D, Arora, L and Mukhopadhyay, S (2021) Fluorescence depolarization kinetics captures short-range backbone dihedral rotations and long-range correlated dynamics of an intrinsically disordered protein. The Journal of Physical Chemistry B 125, 97089718.CrossRefGoogle ScholarPubMed
Davey, NE, Van Roey, K, Weatheritt, RJ, Toedt, G, Uyar, B, Altenberg, B, Budd, A, Diella, F, Dinkel, H and Gibson, TJ (2012) Attributes of short linear motifs. Molecular BioSystems 8, 268281.CrossRefGoogle ScholarPubMed
Deranek, AE, Baldo, AP, Lynn, ML, Schwartz, SD and Tardiff, JC (2022) Structure and dynamics of the flexible cardiac troponin T linker domain in a fully reconstituted thin filament. Biochemistry 61, 12291242.CrossRefGoogle Scholar
Deribe, YL, Pawson, T and Dikic, I (2010) Post-translational modifications in signal integration. Nature Structural & Molecular Biology 17, 666672.CrossRefGoogle ScholarPubMed
DeSantiago, J, Maier, LS and Bers, DM (2002) Frequency-dependent acceleration of relaxation in the heart depends on CaMKII, but not phospholamban. Journal of Molecular and Cellular Cardiology 34, 975984.CrossRefGoogle Scholar
de Souza, JV, Zariquiey, FS and Bronowska, AK (2020) Development of charge-augmented three-point water model (CAIPi3P) for accurate simulations of intrinsically disordered proteins. International Journal of Molecular Sciences 21, 6166.CrossRefGoogle ScholarPubMed
Despond, EA and Dawson, JF (2018) Classifying cardiac actin mutations associated with hypertrophic cardiomyopathy. Frontiers in Physiology 9, 405.CrossRefGoogle ScholarPubMed
de Tombe, PP (2003) Cardiac myofilaments: mechanics and regulation. Journal of Biomechanics 36, 721730.CrossRefGoogle ScholarPubMed
de Winter, JM, Molenaar, JP, Yuen, M, van der Pijl, R, Shen, S, Conijn, S, van de Locht, M, Willigenburg, M, Bogaards, SJ, van Kleef, ES, Lassche, S, Persson, M, Rassier, DE, Sztal, TE, Ruparelia, AA, Oorschot, V, Ramm, G, Hall, TE, Xiong, Z, Johnson, CN, Li, F, Kiss, B, Lozano-Vidal, N, Boon, RA, Marabita, M, Nogara, L, Blaauw, B, Rodenburg, RJ, Küsters, B, Doorduin, J, Beggs, AH, Granzier, H, Campbell, K, Ma, W, Irving, T, Malfatti, E, Romero, NB, Bryson-Richardson, RJ, van Engelen, BG, Voermans, NC and Ottenheijm, CA (2020) KBTBD13 is an actin-binding protein that modulates muscle kinetics. The Journal of Clinical Investigation 130, 754767.CrossRefGoogle ScholarPubMed
De, A, Jasani, A, Arora, R and Gambhir, S (2013) Evolution of BRET biosensors from live cell to tissue-scale in vivo imaging. Frontiers in Endocrinology 4, 131.CrossRefGoogle ScholarPubMed
Dimauro, I, Antonioni, A, Mercatelli, N and Caporossi, D (2018) The role of αB-crystallin in skeletal and cardiac muscle tissues. Cell Stress and Chaperones 23, 491505.CrossRefGoogle ScholarPubMed
Ding, C, Wang, S and Zhang, Z (2021) Integrating an enhanced sampling method and small-angle X-ray scattering to study intrinsically disordered proteins. Frontiers in Pharmacology 8, 103.Google ScholarPubMed
Dogan, J, Jonasson, J, Andersson, E, and Jemth, P (2015) Binding rate constants reveal distinct features of disordered protein domains. Biochemistry 54, 47414750.CrossRefGoogle ScholarPubMed
Doh, CY, Bharambe, N, Holmes, JB, Dominic, KL, Swanberg, CE, Mamidi, R, Chen, Y, Bandyopadhyay, S, Ramachandran, R and Stelzer, JE (2022) Molecular characterization of linker and loop-mediated structural modulation and hinge motion in the C4-C5 domains of cMyBPC. Journal of Structural Biology 214, 107856.CrossRefGoogle ScholarPubMed
Doran, MH, Pavadai, E, Rynkiewicz, MJ, Walklate, J, Bullitt, E, Moore, JR, Regnier, M, Geeves, MA and Lehman, W (2020) Cryo-EM and molecular docking shows myosin loop 4 contacts actin and tropomyosin on thin filaments. Biophysical Journal 119, 821830.CrossRefGoogle ScholarPubMed
Dorovkov, MV, Beznosov, SN, Shah, S, Kotlyanskaya, L and Kostyukova, AS (2008) Effect of mutations imitating the phosphorylation by TRPM7 kinase on the function of the N-terminal domain of tropomodulin. Biophysics 53, 500504.CrossRefGoogle ScholarPubMed
Doucet, D, Roitberg, A and Hagen, SJ (2007) Kinetics of internal-loop formation in polypeptide chains: a simulation study. Biophysical Journal 92, 22812289.CrossRefGoogle ScholarPubMed
Duan, Y, DeKeyser, JG, Damodaran, S and Greaser, ML (2006) Studies on titin PEVK peptides and their interaction. Archives of Biochemistry and Biophysics 454, 1625.CrossRefGoogle ScholarPubMed
Duarte-Costa, S, Castro-Ferreira, R, Neves, JS and Leite-Moreira, AF (2014) S100A1: a major player in cardiovascular performance. Physiological Research 63, 669681.CrossRefGoogle Scholar
Duboscq-Bidot, L, Xu, P, Charron, P, Neyroud, N, Dilanian, G, Millaire, A, Bors, V, Komajda, M and Villard, E (2007) Mutations in the Z-band protein myopalladin gene and idiopathic dilated cardiomyopathy. Cardiovascular Research 77, 118125.CrossRefGoogle ScholarPubMed
Duggal, D, Requena, S, Nagwekar, J, Raut, S, Rich, R, Das, H, Patel, V, Gryczynski, I, Fudala, R, Gryczynski, Z, Blair, C, Campbell, KS and Borejdo, J (2017) No difference in myosin kinetics and spatial distribution of the lever arm in the left and right ventricles of human hearts. Frontiers in Physiology 8, 732.CrossRefGoogle ScholarPubMed
Dvoretsky, A, Abusamhadneh, EM, Howarth, JW and Rosevear, PR (2002) Solution structure of calcium-saturated cardiac troponin C bound to cardiac troponin I. Journal of Biological Chemistry 277, 3856538570.CrossRefGoogle ScholarPubMed
Dyson, HJ and Wright, PE (2021) NMR illuminates intrinsic disorder. Current Opinion in Structural Biology 70, 4452.CrossRefGoogle ScholarPubMed
Earl, DJ and Deem, MW (2008) Monte Carlo simulations. Methods in Molecular Biology 443, 2536.CrossRefGoogle ScholarPubMed
Elkins, JM, Gileadi, C, Shrestha, L, Phillips, C, Wang, J, Muniz, JRC and Doyle, DA (2010) Unusual binding interactions in PDZ domain crystal structures help explain binding mechanisms. Protein Science 19, 731741.CrossRefGoogle ScholarPubMed
Erlendsson, S and Teilum, K (2021) Binding revisited – avidity in cellular function and signaling. Frontiers in Molecular Biosciences 7, 615565.CrossRefGoogle ScholarPubMed
Espinoza-Fonseca, LM, Kast, D and Thomas, DD (2007) Molecular dynamics simulations reveal a disorder-to-order transition on phosphorylation of smooth muscle myosin. Biophysical Journal 93, 20832090.CrossRefGoogle ScholarPubMed
Fanning, SW, Jeselsohn, R, Dharmarajan, V, Mayne, CG, Karimi, M, Buchwalter, G, Houtman, R, Toy, W, Fowler, CE, Han, R, Lainé, M, Carlson, KE, Martin, TA, Nowak, J, Nwachukwu, JC, Hosfield, DJ, Chandarlapaty, S, Tajkhorshid, E, Nettles, KW, Griffin, PR, Shen, Y, Katzenellenbogen, JA, Brown, M and Greene, GL (2018) The SERM/SERD bazedoxifene disrupts ESR1 helix 12 to overcome acquired hormone resistance in breast cancer cells. eLife 7, e37161.CrossRefGoogle ScholarPubMed
Fenwick, AJ, Wood, AM and Tanner, BCW (2017) Effects of cross-bridge compliance on the force-velocity relationship and muscle power output. PLoS ONE 12, e0190335.CrossRefGoogle ScholarPubMed
Fisher, CK and Stultz, CM (2011) Constructing ensembles for intrinsically disordered proteins. Current Opinion in Structural Biology 21, 426431.CrossRefGoogle ScholarPubMed
Fisher, CK, Huang, A and Stultz, CM (2010) Modeling intrinsically disordered proteins with Bayesian statistics. Journal of the American Chemical Society 132, 1491914927.CrossRefGoogle ScholarPubMed
Flashman, E, Redwood, C, MoolmanSmook, J and Watkins, H (2004) Cardiac myosin binding protein C. Circulation Research 94, 12791289.CrossRefGoogle ScholarPubMed
Forman, D and Bulwer, BE (2006) Cardiovascular disease: optimal approaches to risk factor modification of diet and lifestyle. Current Treatment Options in Cardiovascular Medicine 8, 4757.CrossRefGoogle ScholarPubMed
Frank, D, Yusuf Rangrez, A, Friedrich, C, Dittmann, S, Stallmeyer, B, Yadav, P, Bernt, A, Schulze-Bahr, E, Borlepawar, A, Zimmermann, W-H, Peischard, S, Seebohm, G, Linke, WA, Baba, HA, Krüger, M, Unger, A, Usinger, P, Frey, N and Schulze-Bahr, E (2019) Cardiac α-actin (ACTC1) gene mutation causes atrial-septal defects associated with late-onset dilated cardiomyopathy. Circulation: Genomic and Precision Medicine 12, e002491.Google ScholarPubMed
Frey, N, Richardson, JA and Olson, EN (2000) Calsarcins, a novel family of sarcomeric calcineurin-binding proteins. Proceedings of the National Academy of Sciences 97, 1463214637.CrossRefGoogle ScholarPubMed
Friedrich, FW, Reischmann, S, Schwalm, A, Unger, A, Ramanujam, D, Münch, J, Müller, OJ, Hengstenberg, C, Galve, E, Charron, P, Linke, WA, Engelhardt, S, Patten, M, Richard, P, van der Velden, J, Eschenhagen, T, Isnard, R and Carrier, L (2014) FHL2 expression and variants in hypertrophic cardiomyopathy. Basic Research in Cardiology 109, 451.CrossRefGoogle ScholarPubMed
Fukuchi, S, Sakamoto, S, Nobe, Y, Murakami, SD, Amemiya, T, Hosoda, K, Koike, R, Hiroaki, H and Ota, M (2012) IDEAL: intrinsically disordered proteins with extensive annotations and literature. Nucleic Acids Research 40, D507D511.CrossRefGoogle ScholarPubMed
Funk, J, Merino, F, Schaks, M, Rottner, K, Raunser, S and Bieling, P (2021) A barbed end interference mechanism reveals how capping protein promotes nucleation in branched actin networks. Nature Communications 12, 5329.CrossRefGoogle ScholarPubMed
Garaizar, A, Sanchez-Burgos, I, Collepardo-Guevara, R and Espinosa, JR (2020) Expansion of intrinsically disordered proteins increases the range of stability of liquid–liquid phase separation. Molecules 25, 4705.CrossRefGoogle ScholarPubMed
Geisler, SB, Robinson, D, Hauringa, M, Raeker, MO, Borisov, AB, Westfall, MV and Russell, MW (2007) Obscurin-like 1, OBSL1, is a novel cytoskeletal protein related to obscurin. Genomics 89, 521531.CrossRefGoogle ScholarPubMed
Ghisaidoobe, ABT and Chung, SJ (2014) Intrinsic tryptophan fluorescence in the detection and analysis of proteins: a focus on Förster resonance energy transfer techniques. International Journal of Molecular Sciences 15, 2251822538.CrossRefGoogle ScholarPubMed
Ghosh, K, Huihui, J, Phillips, M and Haider, A (2022) Rules of physical mathematics govern intrinsically disordered proteins. Annual Review of Biophysics 51, 355376.CrossRefGoogle ScholarPubMed
Gibbs, EB and Showalter, SA (2015) Quantitative biophysical characterization of intrinsically disordered proteins. Biochemistry 54, 13141326.CrossRefGoogle ScholarPubMed
Gil Pineda, LI, Milko, LN and He, Y (2020) Performance of CHARMM36m with modified water model in simulating intrinsically disordered proteins: a case study. Biophysics Reports 6, 8087.CrossRefGoogle Scholar
Goldfarb, LG, Park, KY, Cervenáková, L, Gorokhova, S, Lee, HS, Vasconcelos, O, Nagle, JW, Semino-Mora, C, Sivakumar, K and Dalakas, MC (1998) Missense mutations in desmin associated with familial cardiac and skeletal myopathy. Nature Genetics 19, 402403.CrossRefGoogle ScholarPubMed
Goldfarb, LG, Olivé, M, Vicart, P and Goebel, HH (2008) Intermediate filament diseases: desminopathy. In Laing, NG (ed.), The Sarcomere and Skeletal Muscle Disease. New York, NY: Springer New York, pp. 131164.CrossRefGoogle Scholar
Golenhofen, N, Arbeiter, A, Koob, R and Drenckhahn, D (2002) Ischemia-induced association of the stress protein B-crystallin with I-band portion of cardiac titin. Journal of Molecular and Cellular Cardiology 34, 309319.CrossRefGoogle ScholarPubMed
Gong, X, Zhang, Y and Chen, J (2021) Advanced sampling methods for multiscale simulation of disordered proteins and dynamic interactions. Biomolecules 11, 1416.CrossRefGoogle ScholarPubMed
Grantham, J (2020) The molecular chaperone CCT/TRiC: an essential component of proteostasis and a potential modulator of protein aggregation. Frontiers in Genetics 11, 172.CrossRefGoogle Scholar
Granzier, H, Fukushima, H and Chung, CS (2010) Titin-isoform dependence of titin-actin interaction and its regulation by S100A1/ Ca2+ in skinned myocardium. Journal of Biomedicine and Biotechnology 2010, 727239.Google Scholar
Greenfield, NJ, Kostyukova, AS and Hitchcock-DeGregori, SE (2005) Structure and tropomyosin binding properties of the N-terminal capping domain of tropomodulin 1. Biophysical Journal 88, 372383.CrossRefGoogle ScholarPubMed
Grotz, KK and Schwierz, N (2022) Optimized magnesium force field parameters for biomolecular simulations with accurate solvation, ion-binding, and water-exchange properties in SPC/E, TIP3P-fb, TIP4P/2005, TIP4P-Ew, and TIP4P-D. Journal of Chemical Theory and Computation 18, 526537.CrossRefGoogle ScholarPubMed
Grupi, A and Haas, E (2011) Segmental conformational disorder and dynamics in the intrinsically disordered protein synuclein and its chain length dependence. Journal of Molecular Biology 405, 12671283.CrossRefGoogle ScholarPubMed
Gupta, A, Dey, S, Hicks, A and Zhou, H-X (2022) Artificial intelligence guided conformational mining of intrinsically disordered proteins. Communications Biology 5, 610.CrossRefGoogle ScholarPubMed
Gurel, PS, Kim, LY, Ruijgrok, PV, Omabegho, T, Bryant, Z and Alushin, GM (2017) Cryo-EM structures reveal specialization at the myosin VI-actin interface and a mechanism of force sensitivity. eLife 6, e31125.CrossRefGoogle Scholar
Hadzi, S, Loris, R and Lah, J (2021) The sequence-ensemble relationship in fuzzy protein complexes. Proceedings of the National Academy of Sciences 118, e2020562118.CrossRefGoogle ScholarPubMed
Hamdani, N, de Waard, M, Messer, AE, Boontje, NM, Kooij, V, van Dijk, S, Versteilen, A, Lamberts, R, Merkus, D, Remedios, CD, Duncker, DJ, Borbely, A, Papp, Z, Paulus, W, Stienen, GJM, Marston, SB and van de Velden, J (2008) Myofilament dysfunction in cardiac disease from mice to men. Journal of Muscle Research and Cell Motility 29, 189201.CrossRefGoogle Scholar
Hamelberg, D, Mongan, J and McCammon, JA (2004) Accelerated molecular dynamics: a promising and efficient simulation method for biomolecules. The Journal of Chemical Physics 120, 1191911929.CrossRefGoogle ScholarPubMed
Hamelberg, D, Shen, T and Andrew McCammon, J (2005) Relating kinetic rates and local energetic roughness by accelerated molecular-dynamics simulations. The Journal of Chemical Physics 122, 241103.CrossRefGoogle ScholarPubMed
Harmon, TS, Holehouse, AS, Rosen, MK and Pappu, RV (2017) Intrinsically disordered linkers determine the interplay between phase separation and gelation in multivalent proteins. eLife 6, e30294.CrossRefGoogle ScholarPubMed
Harrigan, MP, Sultan, MM, Hernández, CX, Husic, BE, Eastman, P, Schwantes, CR, Beauchamp, KA, McGibbon, RT and Pande, VS (2017) MSMBuilder: statistical models for biomolecular dynamics. Biophysical Journal 112, 1015.CrossRefGoogle ScholarPubMed
Harris, SP, Lyons, RG and Bezold, KL (2011) In the thick of it: hCM-causing mutations in myosin binding proteins of the thick filament. Circulation Research 108, 751764.CrossRefGoogle ScholarPubMed
Hartman, MA and Spudich, JA (2012) The myosin superfamily at a glance. Journal of Cell Science 125(Pt 7), 16271632.CrossRefGoogle ScholarPubMed
Heling, LWHJ, Geeves, MA and Kad, NM (2020) MyBP-C: one protein to govern them all. Journal of Muscle Research and Cell Motility 41, 91101.CrossRefGoogle Scholar
Hempel, A and Kuhl, SJ (2014) Comparative expression analysis of cysteine-rich intestinal protein family members crip1, 2 and 3 during Xenopus laevis embryogenesis. The International Journal of Developmental Biology 58, 841849.CrossRefGoogle ScholarPubMed
Henriques, J, Cragnell, C and Skepo, M (2015) Molecular dynamics simulations of intrinsically disordered proteins: force field evaluation and comparison with experiment. Journal of Chemical Theory and Computation 11, 34203431.CrossRefGoogle ScholarPubMed
Herman, DS, Lam, L, Taylor, MR, Wang, L, Teekakirikul, P, Christodoulou, D, Conner, L, DePalma, SR, McDonough, B, Sparks, E, Teodorescu, DL, Cirino, AL, Banner, NR, Pennell, DJ, Graw, S, Merlo, M, Di Lenarda, A, Sinagra, G, Bos, JM, Ackerman, MJ, Mitchell, RN, Murry, CE, Lakdawala, NK, Ho, CY, Barton, PJ, Cook, SA, Mestroni, L, Seidman, J and Seidman, CE (2012) Truncations of titin causing dilated cardiomyopathy. New England Journal of Medicine 366, 619628.CrossRefGoogle ScholarPubMed
Hernandez, DA, Bennett, CM, Dunina-Barkovskaya, L, Wedig, T, Capetanaki, Y, Herrmann, H and Conover, GM (2016) Nebulette is a powerful cytolinker organizing desmin and actin in mouse hearts. Molecular Biology of the Cell 27, 38693882.CrossRefGoogle ScholarPubMed
Hidalgo, CG, Chung, CS, Saripalli, C, Methawasin, M, Hutchinson, KR, Tsaprailis, G, Labeit, S, Mattiazzi, A and Granzier, HL (2013) The multifunctional Ca2+/calmodulin-dependent protein kinase II delta (CaMKII) phosphorylates cardiac titin's spring elements. Journal of Molecular and Cellular Cardiology 54, 9097.CrossRefGoogle Scholar
Hinczewski, M, Hansen, Yvon, Dzubiella, J and Netz, RR (2010) How the diffusivity profile reduces the arbitrariness of protein folding free energies. The Journal of Chemical Physics 132, 245103.CrossRefGoogle ScholarPubMed
Hirano, Y, Amano, Y, Yonemura, S and Hakoshima, T (2018) The force-sensing device region of a-catenin is an intrinsically disordered segment in the absence of intramolecular stabilization of the autoinhibitory form. Genes to Cells 23, 370385.CrossRefGoogle Scholar
Hoffman, RM and Sykes, BD (2008) Isoform-specific variation in the intrinsic disorder of troponin I. Proteins: Structure, Function and Genetics 73, 338350.CrossRefGoogle ScholarPubMed
Hoffman, RM, Blumenschein, TM and Sykes, BD (2006) An interplay between protein disorder and structure confers the Ca2+ regulation of striated muscle. Journal of Molecular Biology 361, 625633.CrossRefGoogle ScholarPubMed
Hojayev, B, Rothermel, BA, Gillette, TG and Hill, JA (2012) FHL2 binds calcineurin and represses pathological cardiac growth. Molecular and Cellular Biology 32, 40254034.CrossRefGoogle ScholarPubMed
Holehouse, AS, Das, RK, Ahad, JN, Richardson, MO and Pappu, RV (2017) CIDER: resources to analyze sequence-ensemble relationships of intrinsically disordered proteins. Biophysical Journal 112, 1621.CrossRefGoogle ScholarPubMed
Holmes, WB and Moncman, CL (2008) Nebulette interacts with filamin C. Cell Motility Cytoskeleton 65, 130142.CrossRefGoogle ScholarPubMed
Hopkins, CW, Le Grand, S, Walker, RC and Roitberg, AE (2015) Long-time-step molecular dynamics through hydrogen mass repartitioning. Journal of Chemical Theory and Computation 11, 18641874.CrossRefGoogle ScholarPubMed
Hornbeck, PV, Zhang, B, Murray, B, Kornhauser, JM, Latham, V and Skrzypek, E (2015) PhosphoSitePlus, 2014: mutations, PTMs and recalibrations. Nucleic Acids Research 43, D512D520.CrossRefGoogle ScholarPubMed
Houdusse, A, Kalabokis, VN, Himmel, D, Szent-Gyorgyi, AG and Cohen, C (1999) Atomic structure of scallop myosin subfragment S1 complexed with MgADP: a novel conformation of the myosin head. Cell 97, 459470.CrossRefGoogle ScholarPubMed
Hough, LE, Dutta, K, Sparks, S, Temel, DB, Kamal, A, Tetenbaum-Novatt, J, Rout, MP and Cowburn, D (2015) The molecular mechanism of nuclear transport revealed by atomic-scale measurements. eLife 4, e10027.CrossRefGoogle ScholarPubMed
Howarth, JW, Ramisetti, S, Nolan, K, Sadayappan, S and Rosevear, PR (2012) Structural insight into unique cardiac myosin-binding protein-C motif: a partially folded domain*. Journal of Biological Chemistry 287, 82548262.CrossRefGoogle ScholarPubMed
Hsin, J, Strümpfer, J, Lee, EH and Schulten, K (2011) Molecular origin of the hierarchical elasticity of titin: simulation, experiment, and theory. Annual Review of Biophysics 40, 187203.CrossRefGoogle ScholarPubMed
Huang, F and Nau, WM (2003) A conformational flexibility scale for amino acids in peptides. Angewandte Chemie – International Edition 42, 22692272.CrossRefGoogle ScholarPubMed
Huang, X, Qu, R, Ouyang, J, Zhong, S and Dai, J (2020 a) An overview of the cytoskeleton-associated role of PDLIM5. Frontiers in Physiology 11, 975.CrossRefGoogle ScholarPubMed
Huang, Y, Mao, X, Jaarsveld, RHvan, Shu, L, Terhal, PA, Jia, Z, Xi, H, Peng, Y, Yan, H, Yuan, S, Li, Q, Wang, H and Bellen, HJ (2020 b) Variants in CAPZA2, a member of an F-actin capping complex, cause intellectual disability and developmental delay. Human Molecular Genetics 29, 15371546.CrossRefGoogle ScholarPubMed
Huber, GA and McCammon, JA (2010) Browndye: a software package for Brownian dynamics. Computer Physics Communications 181, 18961905.CrossRefGoogle ScholarPubMed
Huke, S and Bers, D (2007) Temporal dissociation of frequency-dependent acceleration of relaxation and protein phosphorylation by CaMKII. Journal of Molecular and Cellular Cardiology 42, 590599.CrossRefGoogle ScholarPubMed
Hummer, G (2005) Position-dependent diffusion coefficients and free energies from Bayesian analysis of equilibrium and replica molecular dynamics simulations. New Journal of Physics 7, 34.CrossRefGoogle Scholar
Husic, BE and Pande, VS (2018) Markov state models: from an art to a science. Journal of the American Chemical Society 140, 23862396.CrossRefGoogle ScholarPubMed
Hwang, PM, Cai, F, Pineda-Sanabria, SE, Corson, DC and Sykes, BD (2014) The cardiac-specific n-terminal region of troponin I positions the regulatory domain of troponin C. Proceedings of the National Academy of Sciences 111, 1441214417.CrossRefGoogle ScholarPubMed
Hyeon, C and Thirumalai, D (2006) Kinetics of interior loop formation in semiflexible chains. The Journal of Chemical Physics 124, 104905.CrossRefGoogle ScholarPubMed
Iconaru, LI, Ban, D, Bharatham, K, Ramanathan, A, Zhang, W, Shelat, AA, Zuo, J and Kriwacki, RW (2015) Discovery of small molecules that inhibit the disordered protein, p27Kip1. Scientific Reports 5, 15686.CrossRefGoogle Scholar
Irving, TC, Konhilas, J, Perry, D, Fischetti, R and de Tombe, PP (2000) Myofilament lattice spacing as a function of sarcomere length in isolated rat myocardium. American Journal of Physiology-Heart and Circulatory Physiology 279, H2568H2573.CrossRefGoogle ScholarPubMed
Ivarsson, Y (2012) Plasticity of PDZ domains in ligand recognition and signaling. FEBS Letters 586, 26382647.CrossRefGoogle ScholarPubMed
Izadi, S, Anandakrishnan, R and Onufriev, AV (2014) Building water models: a different approach. The Journal of Physical Chemistry Letters 5, 38633871.CrossRefGoogle ScholarPubMed
Jehle, S, Rajagopal, P, Bardiaux, B, Markovic, S, Kühne, R, Stout, JR, Higman, VA, Klevit, RE, van Rossum, B-J and Oschkinat, H (2010) Solid-state NMR and SAXS studies provide a structural basis for the activation of alphaB-crystallin oligomers. Nature Structural & Molecular Biology 17, 10371042.CrossRefGoogle ScholarPubMed
Jideama, NM, Crawford, BH, Hussain, AKMA and Raynor, RL (2006) Dephosphorylation specificities of protein phosphatase for cardiac troponin I, troponin T, and sites within troponin T. International Journal of Biological Sciences 2, 19.CrossRefGoogle ScholarPubMed
Jin, SC, Homsy, J, Zaidi, S, Lu, Q, Morton, S, DePalma, SR, Zeng, X, Qi, H, Chang, W, Sierant, MC, Hung, W-C, Haider, S, Zhang, J, Knight, J, Bjornson, RD, Castaldi, C, Tikhonoa, IR, Bilguvar, K, Mane, SM, Sanders, SJ, Mital, S, Russell, MW, Gaynor, JW, Deanfield, J, Giardini, A, Porter, GA Jr, Srivastava, D, Lo, CW, Shen, Y, Watkins, WS, Yandell, M, Yost, HJ, Tristani-Firouzi, M, Newburger, JW, Roberts, AE, Kim, R, Zhao, H, Kaltman, JR, Goldmuntz, E, Chung, WK, Seidman, JG, Gelb, BD, Seidman, CE, Lifton, RP and Brueckner, M (2017) Contribution of rare inherited and de novo variants in 2,871 congenital heart disease probands. Nature Genetics 49, 15931601.CrossRefGoogle ScholarPubMed
Jin, Y, Diffee, GM, Colman, RJ anderson, RM and Ge, Y (2019) Top-down mass spectrometry of sarcomeric protein post-translational modifications from non-human primate skeletal muscle. Journal of the American Society for Mass Spectrometry 30, 24602469.CrossRefGoogle ScholarPubMed
Johnston, JR, Landim-Vieira, M, Marques, MA, Oliveira, GAD, Gonzalez-Martinez, D, Moraes, AH, He, H, Iqbal, A, Wilnai, Y, Birk, E, Zucker, N, Silva, JL, Chase, PB and Pinto, JR (2019) The intrinsically disordered C terminus of troponin T binds to troponin C to modulate myocardial force generation. Journal of Biological Chemistry 294, 2005420069.CrossRefGoogle ScholarPubMed
Julien, O, Mercier, P, Allen, CN, Fisette, O, Ramos, CHI, Lagüe, P, Blumenschein, TMA and Sykes, BD (2011) Is there nascent structure in the intrinsically disordered region of troponin I? Proteins: Structure, Function, and Bioinformatics 79, 12401250.CrossRefGoogle ScholarPubMed
Jumper, JM, Faruk, NF, Freed, KF and Sosnick, TR (2018) Accurate calculation of side chain packing and free energy with applications to protein molecular dynamics. PLoS Computational Biology 14, 125.CrossRefGoogle ScholarPubMed
Jumper, J, Evans, R, Pritzel, A, Green, T, Figurnov, M, Ronneberger, O, Tunyasuvunakool, K, Bates, R, Žídek, A, Potapenko, A, Bridgland, A, Meyer, C, Kohl, SAA, Ballard, AJ, Cowie, A, Romera-Paredes, B, Nikolov, S, Jain, R, Adler, J, Back, T, Petersen, S, Reiman, D, Clancy, E, Zielinski, M, Steinegger, M, Pacholska, M, Berghammer, T, Bodenstein, S, Silver, D, Vinyals, O, Senior, AW, Kavukcuoglu, K, Kohli, P and Hassabis, D (2021) Highly accurate protein structure prediction with AlphaFold. Nature 596, 583589.CrossRefGoogle ScholarPubMed
Kampourakis, T, Yan, Z, Gautel, M, Sun, Y-B and Irving, M (2014) Myosin binding protein-C activates thin filaments and inhibits thick filaments in heart muscle cells. Proceedings of the National Academy of Sciences 111, 1876318768.CrossRefGoogle ScholarPubMed
Kanchan, K, Fuxreiter, M and Fésüs, L (2015) Physiological, pathological, and structural implications of non-enzymatic protein-protein interactions of the multifunctional human transglutaminase 2. Cellular and Molecular Life Sciences 72, 30093035.CrossRefGoogle ScholarPubMed
Kang, SJ, Shin, KS, Song, WK, Ha, DB, Chung, CH and Kang, MS (1995) Involvement of transglutaminase in myofibril assembly of chick embryonic myoblasts in culture. Journal of Cell Biology 130, 11271136.CrossRefGoogle ScholarPubMed
Kang, H, Pincus, PA, Hyeon, C and Thirumalai, D (2015) Effects of macromolecular crowding on the collapse of biopolymers. Physical Review Letters 114, 068303.CrossRefGoogle ScholarPubMed
Karczewski, KJ, Francioli, LC, Tiao, G, Cummings, BB, Alföldi, J, Wang, Q, Collins, RL, Laricchia, KM, Ganna, A, Birnbaum, DP, Gauthier, LD, Brand, H, Solomonson, M, Watts, NA, Rhodes, D, Singer-Berk, M, England, EM, Seaby, EG, Kosmicki, JA, Walters, RK, Tashman, K, Farjoun, Y, Banks, E, Poterba, T, Wang, A, Seed, C, Whiffin, N, Chong, JX, Samocha, KE, Pierce-Hoffman, E, Zappala, Z, O'Donnell-Luria, AH, Minikel, EV, Weisburd, B, Lek, M, Ware, JS, Vittal, C, Armean, IM, Bergelson, L, Cibulskis, K, Connolly, KM, Covarrubias, M, Donnelly, S, Ferriera, S, Gabriel, S, Gentry, J, Gupta, N, Jeandet, T, Kaplan, D, Llanwarne, C, Munshi, R, Novod, S, Petrillo, N, Roazen, D, Ruano-Rubio, V, Saltzman, A, Schleicher, M, Soto, J, Tibbetts, K, Tolonen, C, Wade, G, Talkowski, ME; Genome Aggregation Database Consortium; Neale, BM, Daly, MJ and MacArthur, DG (2020) The mutational constraint spectrum quantified from variation in 141,456 humans. Nature 581, 434443.CrossRefGoogle ScholarPubMed
Karsai, RPD, Kellermayer, M and Harris, S (2011) Mechanical unfolding of cardiac myosin binding protein-C by atomic force microscopy. Biophysical Journal 101, 19681977.CrossRefGoogle ScholarPubMed
Ka, S, Rao, ST, Pyzalska, D, Drendel, W, Greaser, M and Sundaralingam, M (1988) Refined structure of chicken skeletal muscle troponin C in the two-calcium state at 2-A resolution. Journal of Biological Chemistry 263, 16281647.Google Scholar
Kekenes-Huskey, PM, Scott, CE and Atalay, S (2016) Quantifying the influence of the crowded cytoplasm on small molecule diffusion. The Journal of Physical Chemistry B 120, 86968706.CrossRefGoogle ScholarPubMed
Kekenes-Huskey, PM, Gillette, A, Hake, J and McCammon, JA (2012) Finite element estimation of protein-ligand association rates with post-encounter effects: applications to calcium binding in troponin C and SERCA. Computational Science & Discovery 5, 020.CrossRefGoogle ScholarPubMed
Kekenes-Huskey, PM, Liao, T, Gillette, AK, Hake, JE, Zhang, Y, Michailova, AP, McCulloch, AD and McCammon, JA (2013) Molecular and subcellular-scale modeling of nucleotide diffusion in the cardiac myofilament lattice. Biophysical Journal 105, 21302140.CrossRefGoogle ScholarPubMed
Kekenes-Huskey, PM, Burgess, DE, Sun, B, Bartos, DC, Rozmus, ER anderson, CL, January, CT, Eckhardt, LL and Delisle, BP (2022) Mutation-specific differences in Kv7.1 (KCNQ1) and Kv11.1 (KCNH2) channel dysfunction and long QT syndrome phenotypes. International Journal of Molecular Sciences 23, 7389.CrossRefGoogle ScholarPubMed
Kelly, MA, Caleshu, C, Morales, A, Buchan, J, Wolf, Z, Harrison, SM, Cook, S, Dillon, MW, Garcia, J, Haverfield, E, Jongbloed, JDH, Macaya, D, Manrai, A, Orland, K, Richard, G, Spoonamore, K, Thomas, M, Thomson, K, Vincent, LM, Walsh, R, Watkins, H, Whiffin, N, Ingles, J, Tintelen, JPvan, Semsarian, C, Ware, JS, Hershberger, R, Funke, B and Cardiovascular Clinical Domain Working Group for theClinGen (2018) Adaptation and validation of the ACMG/AMP variant classification framework for MYH7-associated inherited cardiomyopathies: recommendations by ClinGen's Inherited Cardiomyopathy Expert Panel. Genetics in Medicine 20, 351359.CrossRefGoogle ScholarPubMed
Kelly, C, Pace, N, Gage, M and Pfuhl, M (2021) Solution NMR structure of titin N2A region Ig domain I83 and its interaction with metal ions. Journal of Molecular Biology 433, 166977.CrossRefGoogle ScholarPubMed
Khaymina, SS, Kenney, JM, Schroeter, MM and Chalovich, JM (2007) Fesselin is a natively unfolded protein. Journal of Proteome Research 6, 36483654.CrossRefGoogle ScholarPubMed
Kim, JS and Yethiraj, A (2010) Crowding effects on association reactions at membranes. Biophysical Journal 98, 951958.CrossRefGoogle ScholarPubMed
Kim, J-Y, Meng, F, Yoo, J and Chung, HS (2018) Diffusion-limited association of disordered protein by non-native electrostatic interactions. Nature Communications 9, 4707.CrossRefGoogle ScholarPubMed
Kim, HY, Park, JE, Lee, S-C, Jeon, E-S, On, YK, Kim, SM, Choe, YH, Ki, C-S, Kim, J-W and Kim, KH (2020) Genotype-related clinical characteristics and myocardial fibrosis and their association with prognosis in hypertrophic cardiomyopathy. Journal of Clinical Medicine 9, 1671.CrossRefGoogle ScholarPubMed
Knöll, R, Hoshijima, M, Hoffman, HM, Person, V, Lorenzen-Schmidt, I, Bang, M-L, Hayashi, T, Shiga, N, Yasukawa, H, Schaper, W, McKenna, W, Yokoyama, M, Schork, NJ, Omens, JH, McCulloch, AD, Kimura, A, Gregorio, CC, Poller, W, Schaper, J, Schultheiss, HP and Chien, KR (2002) The cardiac mechanical stretch sensor machinery involves a Z disc complex that is defective in a subset of human dilated cardiomyopathy. Cell 111, 943955.CrossRefGoogle Scholar
Kooij, V, Venkatraman, V, Kirk, JA, Ubaida-Mohien, C, Graham, DR, Faber, MJ and Van Eyk, JE (2014) Identification of cardiac myofilament protein isoforms using multiple mass spectrometry based approaches. PROTEOMICS Clinical Applications 8, 578589.CrossRefGoogle ScholarPubMed
Kostan, J, Pavšič, M, Puž, V, Schwarz, TC, Drepper, F, Molt, S, Graewert, MA, Schreiner, C, Sajko, S, Ven, PFMvan der, Onipe, A, Svergun, DI, Warscheid, B, Konrat, R, Fürst, DO, Lenarčič, B and Djinović-Carugo, K (2021) Molecular basis of F-actin regulation and sarcomere assembly via myotilin. PLoS Biology 19, e3001148.CrossRefGoogle ScholarPubMed
Kosta, S, Colli, D, Ye, Q and Campbell, KS (2022) FiberSim: a flexible open-source model of myofilament-level contraction. Biophysical Journal 121, 175182.CrossRefGoogle ScholarPubMed
Kostyukova, AS, Hitchcock-DeGregori, SE and Greenfield, NJ (2007) Molecular basis of tropomyosin binding to tropomodulin, an actin-capping protein. Journal of Molecular Biology 372, 608618.CrossRefGoogle ScholarPubMed
Kostyukova, A, Maeda, K, Yamauchi, E, Krieger, I and Maeda, Y (2000) Domain structure of tropomodulin. European Journal of Biochemistry 267, 64706475.CrossRefGoogle ScholarPubMed
Krieger, F, Moglich, A and Kiefhaber, T (2005) Effect of proline and glycine residues on dynamics and barriers of loop formation in polypeptide chains. Journal of the American Chemical Society 127, 33463352.CrossRefGoogle ScholarPubMed
Krishnan, N, Koveal, D, Miller, DH, Xue, B, Akshinthala, SD, Kragelj, J, Jensen, MR, Gauss, C-M, Page, R, Blackledge, M, Muthuswamy, SK, Peti, W and Tonks, NK (2014) Targeting the disordered C terminus of PTP1B with an allosteric inhibitor. Nature Chemical Biology 10, 558566.CrossRefGoogle ScholarPubMed
Labeit, S and Kolmerer, B (1995) The complete primary structure of human nebulin and its correlation to muscle structure. Journal of Molecular Biology 248, 308315.CrossRefGoogle ScholarPubMed
Laio, A and Parrinello, M (2002) Escaping free-energy minima. Proceedings of the National Academy of Sciences 99, 1256212566.CrossRefGoogle ScholarPubMed
Lamber, EP, Guicheney, P and Pinotsis, N (2022) The role of the M-band myomesin proteins in muscle integrity and cardiac disease. Journal of Biomedical Science 29, 18.CrossRefGoogle ScholarPubMed
Landrum, MJ, Lee, JM, Riley, GR, Jang, W, Rubinstein, WS, Church, DM and Maglott, DR (2013) ClinVar: public archive of relationships among sequence variation and human phenotype. Nucleic Acids Research 42, D980D985.CrossRefGoogle ScholarPubMed
Lange, S, Ouyang, K, Meyer, G, Cui, L, Cheng, H, Lieber, RL and Chen, J (2009) Obscurin determines the architecture of the longitudinal sarcoplasmic reticulum. Journal of Cell Science 122, 26402650.CrossRefGoogle ScholarPubMed
Latham, AP and Zhang, B (2021) Consistent force field captures homologue-resolved HP1 phase separation. Journal of Chemical Theory and Computation 17, 31343144.CrossRefGoogle ScholarPubMed
Lau, E, Han, Y, Williams, DR, Thomas, CT, Shrestha, R, Wu, JC and Lam, MP (2019) Splice-junction-based mapping of alternative isoforms in the human proteome. Cell Reports 29, 37513765.e5.CrossRefGoogle ScholarPubMed
Lazar, T, Martnez-Prez, E, Quaglia, F, Hatos, A, Chemes, LB, Iserte, JA, Mndez, NA, Garrone, NA, Saldao, TE, Marchetti, J, Rueda, AJV, Bernad, P, Blackledge, M, Cordeiro, TN, Fagerberg, E, Forman-Kay, JD, Fornasari, MS, Gibson, TJ, Gomes, GNW, Gradinaru, CC, Head-Gordon, T, Jensen, MR, Lemke, EA, Longhi, S, Marino-Buslje, C, Minervini, G, Mittag, T, Monzon, AM, Pappu, RV, Parisi, G, Ricard-Blum, S, Ruff, KM, Salladini, E, Skep, M, Svergun, D, Vallet, SD, Varadi, M, Tompa, P, Tosatto, SC and Piovesan, D (2021) PED in 2021: a major update of the protein ensemble database for intrinsically disordered proteins. Nucleic Acids Research 49, D404D411.CrossRefGoogle Scholar
Lee, T, Moran-Gutierrez, CR and Deniz, AA (2015) Probing protein disorder and complexity at single-molecule resolution. Seminars in Cell & Developmental Biology 37, 2634.CrossRefGoogle ScholarPubMed
Lee, EH, Hsin, J, Mayans, O and Schulten, K (2007) Secondary and tertiary structure elasticity of titin Z1Z2 and a titin chain model. Biophysical Journal 93, 17191735.CrossRefGoogle Scholar
Lehman, W (2016) Thin filament structure and the steric blocking model. Comprehensive Physiology 6, 10431069.CrossRefGoogle ScholarPubMed
LeWinter, MM (2005) Functional consequences of sarcomeric protein abnormalities in failing myocardium. Heart Failure Reviews 10, 249257.CrossRefGoogle ScholarPubMed
LeWinter, MM and Granzier, H (2010) Cardiac titin. Circulation 121, 21372145.CrossRefGoogle ScholarPubMed
Liddy, KA, White, MY and Cordwell, SJ (2013) Functional decorations: post-translational modifications and heart disease delineated by targeted proteomics. Genome Medicine 5, 20.CrossRefGoogle ScholarPubMed
Lieutaud, P, Ferron, F, Uversky, AV, Kurgan, L, Uversky, VN and Longhi, S (2016) How disordered is my protein and what is its disorder for? A guide through the ‘dark side’ of the protein universe. Intrinsically Disordered Proteins 4, e1259708.CrossRefGoogle ScholarPubMed
Lindert, S, Kekenes-Huskey, PM and McCammon, JA (2012 b) Long-timescale molecular dynamics simulations elucidate the dynamics and kinetics of exposure of the hydrophobic patch in troponin C. Biophysical Journal 103, 17841789.CrossRefGoogle ScholarPubMed
Lindert, S, Kekenes-Huskey, PM, Huber, G, Pierce, L and McCammon, JA (2012 a) Dynamics and calcium association to the N-terminal regulatory domain of human cardiac troponin C: a multiscale computational study. Journal of Physical Chemistry B 116, 84498459.CrossRefGoogle Scholar
Lindert, S, Cheng, Y, Kekenes-Huskey, P, Regnier, M and McCammon, J (2015) Effects of HCM cTnI mutation R145G on troponin structure and modulation by PKA phosphorylation elucidated by molecular dynamics simulations. Biophysical Journal 108, 395407.CrossRefGoogle ScholarPubMed
Lindorff-Larsen, K and Kragelund, BB (2021) On the potential of machine learning to examine the relationship between sequence, structure, dynamics and function of intrinsically disordered proteins. Journal of Molecular Biology 433, 167196.CrossRefGoogle ScholarPubMed
Lindorff-Larsen, K, Piana, S, Palmo, K, Maragakis, P, Klepeis, JL, Dror, RO and Shaw, DE (2010) Improved side-chain torsion potentials for the Amber ff99SB protein force field. Proteins: Structure, Function, and Bioinformatics 78, 19501958.CrossRefGoogle ScholarPubMed
Linke, WA, Ivemeyer, M, Mundel, P, Stockmeier, MR and Kolmerer, B (1998) Nature of PEVK-titin elasticity in skeletal muscle. Proceedings of the National Academy of Sciences 95, 80528057.CrossRefGoogle ScholarPubMed
Linke, WA, Kulke, M, Li, H, Fujita-Becker, S, Neagoe, C, Manstein, DJ, Gautel, M and Fernandez, JM (2002) PEVK domain of titin: an entropic spring with actin-binding properties. Journal of Structural Biology 137, 194205.CrossRefGoogle ScholarPubMed
Linnemann, A, Vakeel, P, Bezerra, E, Orfanos, Z, Djinović-Carugo, K, Ven, PFMvan der, Kirfel, G and Fürst, DO (2013) Myopodin is an F-actin bundling protein with multiple independent actin-binding regions. Journal of Muscle Research and Cell Motility 34, 6169.CrossRefGoogle ScholarPubMed
Lin, YH, Schmidt, W, Fritz, KS, Jeong, MY, Cammarato, A, Foster, DB, Biesiadecki, J, McKinsey, TA and Woulfe, KC (2020) Site-specific acetyl-mimetic modification of cardiac troponin I modulates myofilament relaxation and calcium sensitivity. Journal of Molecular and Cellular Cardiology 139, 135147.CrossRefGoogle ScholarPubMed
Lipari, G, Szabo, A and Levy, RM (1982) Protein dynamics and NMR relaxation: comparison of simulations with experiment. Nature 300, 197198.CrossRefGoogle Scholar
Liu, Y, Wang, X and Liu, B (2019) A comprehensive review and comparison of existing computational methods for intrinsically disordered protein and region prediction. Briefings in Bioinformatics 20, 330346.CrossRefGoogle ScholarPubMed
Liu, P, Kim, B, Friesner, RA and Berne, BJ (2005) Replica exchange with solute tempering: a method for sampling biological systems in explicit water. Proceedings of the National Academy of Sciences 102, 1374913754.CrossRefGoogle ScholarPubMed
Liu, P, Shi, Q, Daumé, H and Voth, GA (2008) A Bayesian statistics approach to multiscale coarse graining. The Journal of Chemical Physics 129, 214114.CrossRefGoogle ScholarPubMed
Liu, F, Chu, X, Lu, HP and Wang, J (2017) Molecular mechanism of multispecific recognition of Calmodulin through conformational changes. Proceedings of the National Academy of Sciences 114, E3927E3934.Google ScholarPubMed
Li, X, Romero, P, Rani, M, Dunker, AK and Obradovic, Z (1999) Predicting protein disorder for N-, C-, and internal regions. Genome informatics. Workshop on Genome Informatics 10, 3040.Google Scholar
Li, Y, Zhu, G, Paolocci, N, Zhang, P, Takahashi, C, Okumus, N, Heravi, A, Keceli, G, Ramirez-Correa, G, Kass, DA and Murphy, AM (2017) Heart failure-related hyperphosphorylation in the cardiac troponin I C terminus has divergent effects on cardiac function in vivo. Circulation: Heart Failure 10, e003850.Google ScholarPubMed
Li, L, Casalini, T, Arosio, P and Salvalaglio, M (2022) Modeling the structure and interactions of intrinsically disordered peptides with multiple replica, metadynamics-based sampling methods and force-field combinations. Journal of Chemical Theory and Computation 18, 19151928.CrossRefGoogle ScholarPubMed
Lohanadan, K, Molt, S, Dierck, F, van, der Ven PF, Frey, N, Höhfeld, J and Fürst, DO (2021) Isoform-specific functions of synaptopodin-2 variants in cytoskeleton stabilization and autophagy regulation in muscle under mechanical stress. Experimental Cell Research 408, 112865.CrossRefGoogle ScholarPubMed
Long, F, McElheny, D, Jiang, S, Park, S, Caffrey, MS and Fung, LW-M (2007) Conformational change of erythroid a-spectrin at the tetramerization site upon binding-spectrin. Protein Science 16, 25192530.CrossRefGoogle Scholar
Lorand, L and Graham, RM (2003) Transglutaminases: crosslinking enzymes with pleiotropic functions. Nature Reviews Molecular Cell Biology 4, 140156.CrossRefGoogle ScholarPubMed
Lu, Q-W, Wu, X-Y and Morimoto, S (2013) Inherited cardiomyopathies caused by troponin mutations. Journal of Geriatric Cardiology 10, 91101.Google ScholarPubMed
Lu, H, Isralewitz, B, Krammer, A, Vogel, V and Schulten, K (1998) Unfolding of titin immunoglobulin domains by steered molecular dynamics simulation. Biophysical Journal 75, 662671.CrossRefGoogle ScholarPubMed
Lynch 4th TL, Kumar, M, McNamara, JW, Kuster, DWD, Sivaguru, M, Singh, RR, Previs, MJ, Lee, KH, Kuffel, G, Zilliox, MJ, Lin, BL, Ma, W, Gibson, AM, Blaxall, BC, Nieman, ML, Lorenz, JN, Leichter, DM, Leary, OP, Janssen, PML, Tombe, PPde, Gilbert, RJ, Craig, R, Irving, T, Warshaw, DM and Sadayappan, S (2021) Amino terminus of cardiac myosin binding protein-C regulates cardiac contractility. Journal of Molecular and Cellular Cardiology 156, 3344.Google Scholar
Lynn, ML, Tal, Grinspan L, Holeman, TA, Jimenez, J, Strom, J and Tardiff, JC (2017) The structural basis of alpha-tropomyosin linked (Asp230Asn) familial dilated cardiomyopathy. Journal of Molecular and Cellular Cardiology 108, 127137.CrossRefGoogle ScholarPubMed
Ma, Z and Miao, Y (2020) Review: F-actin remodelling during plant signal transduction via biomolecular assembly. Plant Science 301, 110663.CrossRefGoogle ScholarPubMed
Ma, K and Wang, K (2003) Malleable conformation of the elastic PEVK segment of titin: non-co-operative interconversion of polyproline II helix, beta-turn and unordered structures. The Biochemical Journal 374, 687695.CrossRefGoogle ScholarPubMed
MacKerell, AD, Bashford, D, Bellott, M, Dunbrack, RL, Evanseck, JD, Field, MJ, Fischer, S, Gao, J, Guo, H, Ha, S, Joseph-McCarthy, D, Kuchnir, L, Kuczera, K, Lau, FTK, Mattos, C, Michnick, S, Ngo, T, Nguyen, DT, Prodhom, B, Reiher, WE, Roux, B, Schlenkrich, M, Smith, JC, Stote, R, Straub, J, Watanabe, M, Wiórkiewicz-Kuczera, J, Yin, D and Karplus, M (1998) All-atom empirical potential for molecular modeling and dynamics studies of proteins. The Journal of Physical Chemistry B 102, 35863616.CrossRefGoogle ScholarPubMed
Mahmud, Z, Dhami, PS, Rans, C, Liu, PB and Hwang, PM (2021) Dilated cardiomyopathy mutations and phosphorylation disrupt the active orientation of cardiac troponin C. Journal of Molecular Biology 433, 167010.CrossRefGoogle ScholarPubMed
Maier, LS and Bers, DM (2002) Calcium, calmodulin, and calcium-calmodulin kinase II: heartbeat to heartbeat and beyond. Journal of Molecular and Cellular Cardiology 34, 919939.CrossRefGoogle ScholarPubMed
Main, A, Fuller, W and Baillie, GS (2020) Post-translational regulation of cardiac myosin binding protein-C: a graphical review. Cellular Signalling 76, 109788.CrossRefGoogle ScholarPubMed
Maiweilidan, Y, Klauza, I and Kordeli, E (2011) Novel interactions of ankyrins-G at the costameres: the muscle-specific obscurin/titin-binding-related domain (OTBD) binds plectin and filamin C. Experimental Cell Research 317, 724736.CrossRefGoogle ScholarPubMed
Manukian, S, Punch, E and Gage, M (2022) pH-dependent compaction of intrinsically disordered poly-E motif in titin. Biophysical Journal 121, 200a201a.CrossRefGoogle Scholar
Marques, MA, Parvatiyar, MS, Yang, W, Oliveira, GAPde and Pinto, JR (2019) The missing links within troponin. Archives of Biochemistry and Biophysics 663, 95100.CrossRefGoogle ScholarPubMed
Marrink, SJ, Risselada, HJ, Yefimov, S, Tieleman, DP and de Vries, AH (2007) The MARTINI force field: coarse grained model for biomolecular simulations. The Journal of Physical Chemistry B 111, 78127824.CrossRefGoogle ScholarPubMed
Marsh, JA, Dancheck, B, Ragusa, MJ, Allaire, M, Forman-Kay, JD and Peti, W (2010) Structural diversity in free and bound states of intrinsically disordered protein phosphatase 1 regulators. Structure 18, 10941103.CrossRefGoogle ScholarPubMed
Marston, S (2017) Obscurin variants and inherited cardiomyopathies. Biophysical Reviews 9, 239243.CrossRefGoogle ScholarPubMed
Marston, SB and Redwood, CS (2003) Modulation of thin filament activation by breakdown or isoform switching of thin filament proteins. Circulation Research 93, 11701178.CrossRefGoogle ScholarPubMed
Marston, S and Zamora, JE (2020) Troponin structure and function: a view of recent progress. Journal of Muscle Research and Cell Motility 41, 7189.CrossRefGoogle ScholarPubMed
Martin, TG and Kirk, JA (2020) Under construction: the dynamic assembly, maintenance, and degradation of the cardiac sarcomere. Journal of Molecular and Cellular Cardiology 148, 89102.CrossRefGoogle ScholarPubMed
Martin, EW, Holehouse, AS, Grace, CR, Hughes, A, Pappu, RV and Mittag, T (2016) Sequence determinants of the conformational properties of an intrinsically disordered protein prior to and upon multisite phosphorylation. Journal of the American Chemical Society 138, 1532315335.CrossRefGoogle ScholarPubMed
Maruyama, K and Ebashi, S (1965) Alpha-actinin, a new structural protein from striated muscle. II. Action on actin. The Journal of Biochemistry 58, 1319.CrossRefGoogle ScholarPubMed
Mason, AB, Lynn, ML, Baldo, AP, Deranek, AE, Tardiff, JC and Schwartz, SD (2021) Computational and biophysical determination of pathogenicity of variants of unknown significance in cardiac thin filament. JCI Insight 6, e154350.CrossRefGoogle ScholarPubMed
Maximova, E, Postnikov, EB, Lavrova, AI, Farafonov, V and Nerukh, D (2021) Protein-ligand dissociation rate constant from all-atom simulation. Journal of Physical Chemistry Letters 12, 1063110636.CrossRefGoogle ScholarPubMed
Mazelin, L, Panthu, B, Nicot, A-S, Belotti, E, Tintignac, L, Teixeira, G, Zhang, Q, Risson, V, Baas, D, Delaune, E, Derumeaux, G, Taillandier, D, Ohlmann, T, Ovize, M, Gangloff, Y-G and Schaeffer, L (2016) mTOR inactivation in myocardium from infant mice rapidly leads to dilated cardiomyopathy due to translation defects and p53/JNK-mediated apoptosis. Journal of Molecular and Cellular Cardiology 97, 213225.CrossRefGoogle ScholarPubMed
McLendon, PM and Robbins, J (2011) Desmin-related cardiomyopathy: an unfolding story. American Journal of Physiology Heart and Circulatory Physiology 301, H1220H1228.CrossRefGoogle ScholarPubMed
Meng, F, Uversky, VN and Kurgan, L (2017) Comprehensive review of methods for prediction of intrinsic disorder and its molecular functions. Cellular and Molecular Life Sciences 74, 30693090.CrossRefGoogle ScholarPubMed
Merkley, ED, Cort, JR and Adkins, JN (2013) Cross-linking and mass spectrometry methodologies to facilitate structural biology: finding a path through the maze. Journal of Structural and Functional Genomics 14, 7790.CrossRefGoogle ScholarPubMed
Messer, AE and Marston, SB (2014) Investigating the role of uncoupling of troponin I phosphorylation from changes in myofibrillar Ca(2+)-sensitivity in the pathogenesis of cardiomyopathy. Frontiers in Physiology 5, 315.CrossRefGoogle ScholarPubMed
Mészáros, B, Erdos, G and Dosztányi, Z (2018) IUPred2A: context-dependent prediction of protein disorder as a function of redox state and protein binding. Nucleic Acids Research 46, W329W337.CrossRefGoogle ScholarPubMed
Metskas, LA and Rhoades, E (2015) Conformation and dynamics of the troponin I C-terminal domain: combining single-molecule and computational approaches for a disordered protein region. Journal of the American Chemical Society 137, 1196211969.CrossRefGoogle ScholarPubMed
Metskas, LA and Rhoades, E (2016) Order-disorder transitions in the cardiac troponin complex. Journal of Molecular Biology 428, 29652977.CrossRefGoogle ScholarPubMed
Metskas, LA and Rhoades, E (2020) Single-molecule FRET of intrinsically disordered proteins. Annual Review of Physical Chemistry 71, 391414.CrossRefGoogle ScholarPubMed
Miao, Y, Feher, VA and McCammon, JA (2015) Gaussian accelerated molecular dynamics: unconstrained enhanced sampling and free energy calculation. Journal of Chemical Theory and Computation 11, 35843595.CrossRefGoogle ScholarPubMed
Miao, Y, Tipakornsaowapak, T, Zheng, L, Mu, Y and Lewellyn, E (2018) Phospho-regulation of intrinsically disordered proteins for actin assembly and endocytosis. FEBS Journal 285, 27622784.CrossRefGoogle ScholarPubMed
Michie, K, Kwan, A, Tung, C-S, Guss, J and Trewhella, J (2016) A highly conserved yet flexible linker is part of a polymorphic protein-binding domain in myosin-binding protein C. Structure 24, 20002007.CrossRefGoogle ScholarPubMed
Micsonai, A, Wien, F, Kernya, L, Lee, Y-H, Goto, Y, Réfrégiers, M and Kardos, J (2015) Accurate secondary structure prediction and fold recognition for circular dichroism spectroscopy. Proceedings of the National Academy of Sciences 112, E3095E3103.CrossRefGoogle ScholarPubMed
Milligan, RA, Whittaker, M and Safer, D (1990) Molecular structure of F-actin and location of surface binding sites. Nature 348, 217221.CrossRefGoogle ScholarPubMed
Millman, B and Irving, T (1988) Filament lattice of frog striated muscle. Radial forces, lattice stability, and filament compression in the A-band of relaxed and rigor muscle. Biophysical Journal 54, 437447.CrossRefGoogle ScholarPubMed
Milstein, JN and Meiners, J-C (2013) Worm-like chain (WLC) model. In Roberts, GCK (ed.), Encyclopedia of Biophysics. Berlin, Heidelberg: Springer Berlin Heidelberg, pp. 27572760.CrossRefGoogle Scholar
Mioduszewski, U, Rycki, B and Cieplak, M (2020) Pseudo-improper-dihedral model for intrinsically disordered proteins. Journal of Chemical Theory and Computation 16, 47264733.CrossRefGoogle ScholarPubMed
Misiura, MM and Kolomeisky, AB (2020) Role of intrinsically disordered regions in acceleration of protein–protein association. The Journal of Physical Chemistry B 124, 2027.CrossRefGoogle ScholarPubMed
Moncman, CL and Wang, K (1999) Functional dissection of nebulette demonstrates actin binding of nebulin-like repeats and Z-line targeting of SH3 and linker domains. Cell Motility and the Cytoskeleton 44, 122.3.0.CO;2-8>CrossRefGoogle ScholarPubMed
Moncman, CL and Wang, K (2000) Architecture of the thin filament-Z-line junction: lessons from nebulette and nebulin homologies. Journal of Muscle Research & Cell Motility 21, 153169.CrossRefGoogle ScholarPubMed
Moore, JR, Leinwand, L, Warshaw, DM, Robbins, J, Seidman, C and Watkins, H (2012) Understanding cardiomyopathy phenotypes based on the functional impact of mutations in the myosin motor. Circulation Research 111, 375385.CrossRefGoogle ScholarPubMed
Moroz, N, Guillaud, L, Desai, B and Kostyukova, AS (2013) Mutations changing tropomodulin affinity for tropomyosin alter neurite formation and extension. PeerJ 1, e7.CrossRefGoogle ScholarPubMed
Mostofian, B, McFarland, R, Estelle, A, Howe, J, Barbar, E, Reichow, SL and Zuckerman, DM (2022) Continuum dynamics and statistical correction of compositional heterogeneity in multivalent IDP oligomers resolved by single-particle EM. Journal of Molecular Biology 434, 167520.CrossRefGoogle ScholarPubMed
Moussavi-Baygi, R and Mofrad, MRK (2016) Rapid Brownian motion primes ultrafast reconstruction of intrinsically disordered Phe-Gly repeats inside the nuclear pore complex. Scientific Reports 6, 29991.CrossRefGoogle ScholarPubMed
Mühle, S, Zhou, M, Ghosh, A and Enderlein, J (2019) Loop formation and translational diffusion of intrinsically disordered proteins. Physical Review E 100, 052405.CrossRefGoogle ScholarPubMed
Muranova, LK, Strelkov, SV and Gusev, NB (2020) Effect of cataract-associated mutations in the N-terminal domain of aB-crystallin (HspB5). Experimental Eye Research 197, 108091.CrossRefGoogle Scholar
Mu, J, Liu, H, Zhang, J, Luo, R and Chen, HF (2021) Recent force field strategies for intrinsically disordered proteins. Journal of Chemical Information and Modeling 61, 611037.CrossRefGoogle ScholarPubMed
Nag, S, Trivedi, DV, Sarkar, SS, Adhikari, AS, Sunitha, MS, Sutton, S, Ruppel, KM and Spudich, JA (2017) The myosin mesa and the basis of hypercontractility caused by hypertrophic cardiomyopathy mutations. Nature Structural and Molecular Biology 24, 525533.CrossRefGoogle ScholarPubMed
Nakamura, F, Stossel, TP and Hartwig, JH (2011) The filamins. Cell Adhesion & Migration 5, 160169.CrossRefGoogle ScholarPubMed
Napierski, NC, Granger, K, Langlais, PR, Moran, HR, Strom, J, Touma, K and Harris, SP (2020) A novel ‘cut and paste’ method for in situ replacement of cMyBP-C reveals a new role for cMyBP-C in the regulation of contractile oscillations. Circulation Research 126, 737749.CrossRefGoogle ScholarPubMed
Narita, T, Weinert, BT and Choudhary, C (2019) Functions and mechanisms of non-histone protein acetylation. Nature Reviews Molecular Cell Biology 20, 156174.CrossRefGoogle ScholarPubMed
Na, I, Kong, MJ, Straight, S, Pinto, JR and Uversky, VN (2016) Troponins, intrinsic disorder, and cardiomyopathy. Biological Chemistry 397, 731751.CrossRefGoogle ScholarPubMed
Neirynck, K, Waterschoot, D, Vandekerckhove, J, Ampe, C and Rommelaere, H (2006) Actin interacts with CCT via discrete binding sites: a binding transition-release model for CCT-mediated actin folding. Journal of Molecular Biology 355, 124138.CrossRefGoogle ScholarPubMed
Niederer, SA, Campbell, KS and Campbell, SG (2019) A short history of the development of mathematical models of cardiac mechanics. Journal of Molecular and Cellular Cardiology 127, 1119.CrossRefGoogle Scholar
Northrup, SH, Allison, SA and McCammon, JA (1984) Brownian dynamics simulation of diffusion-influenced bimolecular reactions. The Journal of Chemical Physics 80, 15171524.CrossRefGoogle Scholar
Nott, TJ, Craggs, TD and Baldwin, AJ (2016) Membraneless organelles can melt nucleic acid duplexes and act as biomolecular filters. Nature Chemistry 8, 569575.CrossRefGoogle ScholarPubMed
Oakley, CE, Hambly, BD, Curmi, PM and Brown, LJ (2004) Myosin binding protein C: structural abnormalities in familial hypertrophic cardiomyopathy. Cell Research 14, 95110.CrossRefGoogle ScholarPubMed
Obradovic, Z, Peng, K, Vucetic, S, Radivojac, P, Brown, CJ and Dunker, AK (2003) Predicting intrinsic disorder from amino acid sequence. Proteins: Structure, Function, and Bioinformatics 53(S6), 566572.CrossRefGoogle ScholarPubMed
O'Flynn, BG and Mittag, T (2021) A new phase for enzyme kinetics. Nature Chemical Biology 17, 628630.CrossRefGoogle ScholarPubMed
Okamoto, K and Sako, Y (2017) Recent advances in FRET for the study of protein interactions and dynamics. Current Opinion in Structural Biology 46, 1623.CrossRefGoogle Scholar
Okamoto, R, Li, Y, Noma, K, Hiroi, Y, Liu, P-Y, Taniguchi, M, Ito, M and Liao, JK (2013) FHL2 prevents cardiac hypertrophy in mice with cardiac-specific deletion of ROCK2. FASEB Journal 27, 14391449.CrossRefGoogle ScholarPubMed
Oldfield, CJ, Cheng, Y, Cortese, MS, Romero, P, Uversky, VN and Dunker, AK (2005) Coupled folding and binding with alpha-helix-forming molecular recognition elements. Biochemistry 44, 1245412470.CrossRefGoogle ScholarPubMed
Ono, M, Burgess, DE, Schroder, EA, Elayi, CS, anderson, CL, January, CT, Sun, B, Immadisetty, K, Kekenes-Huskey, PM and Delisle, BP (2020) Long QT syndrome type 2: emerging strategies for correcting class 2 KCNH2 (hERG) mutations and identifying new patients. Biomolecules 10, 1144.CrossRefGoogle ScholarPubMed
Ozenne, V, Bauer, F, Salmon, L, Huang, JR, Jensen, MR, Segard, S, Bernad, P, Charavay, C and Blackledge, M (2012) Flexible-meccano: a tool for the generation of explicit ensemble descriptions of intrinsically disordered proteins and their associated experimental observables. Bioinformatics 28, 14631470.CrossRefGoogle ScholarPubMed
Ozmaian, M and Makarov, DE (2019) Transition path dynamics in the binding of intrinsically disordered proteins: a simulation study. The Journal of Chemical Physics 151, 235101.CrossRefGoogle ScholarPubMed
Pace, T, Rahmaninejad, H, Sun, B and Kekenes-Huskey, PM (2021) Homogenization of continuum-scale transport properties from molecular dynamics simulations: an application to aqueous-phase methane diffusion in silicate channels. The Journal of Physical Chemistry. B 125, 1152011533.CrossRefGoogle ScholarPubMed
Palmer, AG (2004) NMR characterization of the dynamics of biomacromolecules. Chemical Reviews 104, 36233640.CrossRefGoogle ScholarPubMed
Palmer, BM, Sadayappan, S, Wang, Y, Weith, AE, Previs, MJ, Bekyarova, T, Irving, TC, Robbins, J and Maughan, DW (2011) Roles for cardiac MyBP-C in maintaining myofilament lattice rigidity and prolonging myosin cross-bridge lifetime. Biophysical Journal 101, 16611669.CrossRefGoogle ScholarPubMed
Pappas, CT, Mayfield, RM, Henderson, C, Jamilpour, N, Cover, C, Hernandez, Z, Hutchinson, KR, Chu, M, Nam, K-H, Valdez, JM, Wong, PK, Granzier, HL and Gregorio, CC (2015) Knockout of Lmod2 results in shorter thin filaments followed by dilated cardiomyopathy and juvenile lethality. Proceedings of the National Academy of Sciences 112, 1357313578.CrossRefGoogle ScholarPubMed
Parker, F, Batchelor, M, Wolny, M, Hughes, R, Knight, PJ and Peckham, M (2018) A1603P and K1617del, mutations in beta-cardiac myosin heavy chain that cause laing early-onset distal myopathy, affect secondary structure and filament formation in vitro and in vivo. Journal of Molecular Biology 430, 14591478.CrossRefGoogle ScholarPubMed
Park, S, Caffrey, MS, Johnson, ME and Fung, LW-M (2003) Solution structural studies on human erythrocyte α-spectrin tetramerization site. Journal of Biological Chemistry 278, 2183721844.CrossRefGoogle ScholarPubMed
Peng, Y, Gregorich, ZR, Valeja, SG, Zhang, H, Cai, W, Chen, Y-C, Guner, H, Chen, AJ, Schwahn, DJ, Hacker, TA, Liu, X and Ge, Y (2014) Top-down proteomics reveals concerted reductions in myofilament and Z-disc protein phosphorylation after acute myocardial infarction. Molecular & Cellular Proteomics 13, 27522764.CrossRefGoogle ScholarPubMed
Pérez-Hernández, G, Paul, F, Giorgino, T, De Fabritiis, G and Noé, F (2013) Identification of slow molecular order parameters for Markov model construction. The Journal of Chemical Physics 139, 015102.CrossRefGoogle ScholarPubMed
Pfitzer, G (2001) Invited review: regulation of myosin phosphorylation in smooth muscle. Journal of Applied Physiology 91, 497503.CrossRefGoogle ScholarPubMed
Piana, S and Laio, A (2007) A bias-exchange approach to protein folding. The Journal of Physical Chemistry B 111, 45534559.CrossRefGoogle ScholarPubMed
Pietrek, LM, Stelzl, LS and Hummer, G (2020) Hierarchical ensembles of intrinsically disordered proteins at atomic resolution in molecular dynamics simulations. Journal of Chemical Theory and Computation 16, 725737.CrossRefGoogle ScholarPubMed
Pinkas, DM, Strop, P, Brunger, AT and Khosla, C (2007) Transglutaminase 2 undergoes a large conformational change upon activation. PLoS Biology 5, 19.CrossRefGoogle ScholarPubMed
Polizzi, NF, Therien, MJ and Beratan, DN (2016) Mean first-passage times in biology. Israel Journal of Chemistry 56, 816824.CrossRefGoogle ScholarPubMed
Ponder, JW and Case, DA (2003) Force fields for protein simulations. Advances in Protein Chemistry 66, 2785.CrossRefGoogle ScholarPubMed
Potrzebowski, W, Trewhella, J and Andre, I (2018) Bayesian inference of protein conformational ensembles from limited structural data. PLoS Computational Biology 14, e1006641.CrossRefGoogle ScholarPubMed
Povarova, OI, Uversky, VN, Kuznetsova, IM and Turoverov, KK (2014) Actinous enigma or enigmatic actin: folding, structure, and functions of the most abundant eukaryotic protein. Intrinsically Disordered Proteins 2, e34500.CrossRefGoogle ScholarPubMed
Powers, JD, Yuan, C-C, McCabe, KJ, Murray, JD, Childers, MC, Flint, GV, Moussavi-Harami, F, Mohran, S, Castillo, R, Zuzek, C, Ma, W, Daggett, V, McCulloch, AD, Irving, TC and Regnier, M (2019) Cardiac myosin activation with 2-deoxy-ATP via increased electrostatic interactions with actin. Proceedings of the National Academy of Sciences 116, 1150211507.CrossRefGoogle ScholarPubMed
Prabakaran, S, Lippens, G, Steen, H and Gunawardena, J (2012) Post-translational modification: nature's escape from genetic imprisonment and the basis for dynamic information encoding. WIREs Systems Biology and Medicine 4, 565583.CrossRefGoogle ScholarPubMed
Prestegard, JH, Bougault, CM and Kishore, AI (2004) Residual dipolar couplings in structure determination of biomolecules. Chemical Reviews 104, 35193540.CrossRefGoogle ScholarPubMed
Previs, MJ, Mun, JY, Michalek, AJ, Previs, SB, Gulick, J, Robbins, J, Warshaw, DM and Craig, R (2016) Phosphorylation and calcium antagonistically tune myosin-binding protein C's structure and function. Proceedings of the National Academy of Sciences 113, 32393244.CrossRefGoogle Scholar
Prinz, JH, Wu, H, Sarich, M, Keller, B, Senne, M, Held, M, Chodera, JD, Schtte, C and No, F (2011) Markov models of molecular kinetics: generation and validation. The Journal of Chemical Physics 134, 174105.CrossRefGoogle ScholarPubMed
Purevjav, E, Arimura, T, Augustin, S, Huby, A-C, Takagi, K, Nunoda, S, Kearney, DL, Taylor, MD, Terasaki, F, Bos, JM, Ommen, SR, Shibata, H, Takahashi, M, Itoh-Satoh, M, McKenna, WJ, Murphy, RT, Labeit, S, Yamanaka, Y, Machida, N, Park, J-E, Alexander, PMA, Weintraub, RG, Kitaura, Y, Ackerman, MJ, Kimura, A and Towbin, JA (2012) Molecular basis for clinical heterogeneity in inherited cardiomyopathies due to myopalladin mutations. Human Molecular Genetics 21, 20392053.CrossRefGoogle ScholarPubMed
Puž, V, Pavšič, M, Lenarčič, B and Djinović-Carugo, K (2017) Conformational plasticity and evolutionary analysis of the myotilin tandem Ig domains. Scientific Reports 7, 3993.CrossRefGoogle ScholarPubMed
Qiao, Q, Bowman, GR and Huang, X (2013) Dynamics of an intrinsically disordered protein reveal metastable conformations that potentially seed aggregation. Journal of the American Chemical Society 135, 1609216101.CrossRefGoogle ScholarPubMed
Ramirez-Correa, GA, Frazier, AH, Zhu, G, Zhang, P, Rappold, T, Kooij, V, Bedja, D, Snyder, GA, Lugo-Fagundo, NS, Hariharan, R, Li, Y, Shen, X, Gao, WD, Cingolani, OH, Takimoto, E, Foster, DB and Murphy, AM (2015) Cardiac troponin I Pro82Ser variant induces diastolic dysfunction, blunts beta-adrenergic response, and impairs myofilament cooperativity. Journal of Applied Physiology 118, 212223.CrossRefGoogle ScholarPubMed
Ramis, R, Ortega-Castro, J, Casasnovas, R, Marino, L, Vilanova, B, Adrover, M and Frau, J (2019) A coarse-grained molecular dynamics approach to the study of the intrinsically disordered protein α-synuclein. Journal of Chemical Information and Modeling 59, 14581471.CrossRefGoogle Scholar
Rao, V, Cheng, Y, Lindert, S, Wang, D, Oxenford, L, McCulloch, AD, McCammon, JA and Regnier, M (2014) PKA phosphorylation of cardiac troponin I modulates activation and relaxation kinetics of ventricular myofibrils. Biophysical Journal 107, 11961204.CrossRefGoogle ScholarPubMed
Rath, A, Glibowicka, M, Nadeau, VG, Chen, G and Deber, CM (2009) Detergent binding explains anomalous SDS-PAGE migration of membrane proteins. Proceedings of the National Academy of Sciences 106, 17601765.CrossRefGoogle ScholarPubMed
Regy, RM, Thompson, J, Kim, YC and Mittal, J (2021) Improved coarse-grained model for studying sequence dependent phase separation of disordered proteins. Protein Science 30, 13711379.CrossRefGoogle ScholarPubMed
Ribeiro, EDA Jr, Pinotsis, N, Ghisleni, A, Salmazo, A, Konarev, PV, Kostan, J, Sjöblom, B, Schreiner, C, Polyansky, AA, Gkougkoulia, EA, Holt, MR, Aachmann, FL, Zagrović, B, Bordignon, E, Pirker, KF, Svergun, DI, Gautel, M and Djinović-Carugo, K (2014) The structure and regulation of human muscle α-actinin. Cell 159, 14471460.CrossRefGoogle ScholarPubMed
Risi, C, Eisner, J, Belknap, B, Heeley, DH, White, HD, Schröder, GF and Galkin, VE (2017) Ca2+-induced movement of tropomyosin on native cardiac thin filaments revealed by cryoelectron microscopy. Proceedings of the National Academy of Sciences 114, 67826787.CrossRefGoogle Scholar
Robert-Paganin, J, Pylypenko, O, Kikuti, C, Sweeney, HL and Houdusse, A (2020) Force generation by myosin motors: a structural perspective. Chemical Reviews 120, 535.CrossRefGoogle ScholarPubMed
Robustelli, P, Piana, S and Shaw, DE (2018) Developing a molecular dynamics force field for both folded and disordered protein states. Proceedings of the National Academy of Sciences 115, E4758E4766.CrossRefGoogle ScholarPubMed
Robustelli, P, Piana, S, Shaw, DE and Shaw, DE (2020) Mechanism of coupled folding-upon-binding of an intrinsically disordered protein. Journal of the American Chemical Society 142, 1109211101.CrossRefGoogle ScholarPubMed
Romero, P, Obradovic, Z, Li, X, Garner, EC, Brown, CJ and Dunker, AK (2001) Sequence complexity of disordered protein. Proteins 42, 3848.3.0.CO;2-3>CrossRefGoogle ScholarPubMed
Roof, DJ, Hayes, A, Adamian, M, Chishti, AH and Li, T (1997) Molecular characterization of abLIM, a novel actin-binding and double zinc finger protein. Journal of Cell Biology 138, 575588.CrossRefGoogle ScholarPubMed
Ruan, H, Sun, Q, Zhang, W, Liu, Y and Lai, L (2019) Targeting intrinsically disordered proteins at the edge of chaos. Drug Discovery Today 24, 217227.CrossRefGoogle ScholarPubMed
Ruggiero, A, Chen, SN, Lombardi, R, Rodriguez, G and Marian, AJ (2012) Pathogenesis of hypertrophic cardiomyopathy caused by myozenin 2 mutations is independent of calcineurin activity. Cardiovascular Research 97, 4454.CrossRefGoogle ScholarPubMed
Rumi-Masante, J, Rusinga, FI, Lester, TE, Dunlap, TB, Williams, TD, Dunker, AK, Weis, DD and Creamer, TP (2012) Structural basis for activation of calcineurin by calmodulin. Journal of Molecular Biology 415, 307317.CrossRefGoogle ScholarPubMed
Russell, B and Solís, C (2021) Mechanosignaling pathways alter muscle structure and function by post-translational modification of existing sarcomeric proteins to optimize energy usage. Journal of Muscle Research and Cell Motility 42, 367380.CrossRefGoogle ScholarPubMed
Rynkiewicz, MJ, Schott, V, Orzechowski, M, Lehman, W and Fischer, S (2015) Electrostatic interaction map reveals a new binding position for tropomyosin on F-actin. Journal of Muscle Research and Cell Motility 36, 525533.CrossRefGoogle ScholarPubMed
Salmikangas, P, van der Ven, PFM, Lalowski, M, Taivainen, A, Zhao, F, Suila, H, Schröder, R, Lappalainen, P, Fürst, DO and Carpén, O (2003) Myotilin, the limb-girdle muscular dystrophy 1A (LGMD1A) protein, cross-links actin filaments and controls sarcomere assembly. Human Molecular Genetics 12, 189203.CrossRefGoogle ScholarPubMed
Sanbe, A, Fewell, JG, Gulick, J, Osinska, H, Lorenz, J, Hall, DG, Murray, LA, Kimball, TR, Witt, SA and Robbins, J (1999) Abnormal cardiac structure and function in mice expressing nonphosphorylatable cardiac regulatory myosin light chain 2*. Journal of Biological Chemistry 274, 2108521094.CrossRefGoogle ScholarPubMed
Sane, DC, Kontos, JL and Greenberg, CS (2007) Roles of transglutaminases in cardiac and vascular diseases. Frontiers in Bioscience – Landmark 12, 25302545.CrossRefGoogle ScholarPubMed
Santos, J, Iglesias, V, Pintado, C, Santos-Surez, J and Ventura, S (2020) DispHred: a server to predict pH-dependent order disorder transitions in intrinsically disordered proteins. International Journal of Molecular Sciences 21, 5814.CrossRefGoogle ScholarPubMed
Savastano, A, Ibáñez, de Opakua A, Rankovic, M and Zweckstetter, M (2020) Nucleocapsid protein of SARS-CoV-2 phase separates into RNA-rich polymerase-containing condensates. Nature Communications 11, 6041.CrossRefGoogle ScholarPubMed
Scherer, MK, Trendelkamp-Schroer, B, Paul, F, Pérez-Hernández, G, Hoffmann, M, Plattner, N, Wehmeyer, C, Prinz, J-H and Noé, F (2015) PyEMMA 2: a software package for estimation, validation, and analysis of markov models. Journal of Chemical Theory and Computation 11, 55255542.CrossRefGoogle ScholarPubMed
Schneider, R, Blackledge, M and Jensen, MR (2019) Elucidating binding mechanisms and dynamics of intrinsically disordered protein complexes using NMR spectroscopy. Current Opinion in Structural Biology 54, 1018.CrossRefGoogle ScholarPubMed
Schoenauer, R, Bertoncini, P, Machaidze, G, Aebi, U, Perriard, J-C, Hegner, M and Agarkova, I (2005) Myomesin is a molecular spring with adaptable elasticity. Journal of Molecular Biology 349, 367379.CrossRefGoogle ScholarPubMed
Schoenauer, R, Emmert, MY, Felley, A, Ehler, E, Brokopp, C, Weber, B, Nemir, M, Faggian, GG, Pedrazzini, T, Falk, V, Hoerstrup, SP and Agarkova, I (2011) EH-myomesin splice isoform is a novel marker for dilated cardiomyopathy. Basic Research in Cardiology 106, 233247.CrossRefGoogle ScholarPubMed
Schotte, F, Lim, M, Jackson, TA, Smirnov, AV, Soman, J, Olson, JS, Phillips, GN, Wulff, M and Anfinrud, PA (2003) Watching a protein as it functions with 150-ps time-resolved X-ray crystallography. Science 300, 19441947.CrossRefGoogle ScholarPubMed
Schramm, A, Bignon, C, Brocca, S, Grandori, R, Santambrogio, C and Longhi, S (2019) An arsenal of methods for the experimental characterization of intrinsically disordered proteins how to choose and combine them? Archives of Biochemistry and Biophysics 676, 108055.CrossRefGoogle Scholar
Schuler, B, Soranno, A, Hofmann, H and Nettels, D (2016) Single molecule FRET spectroscopy and the polymer physics of unfolded and intrinsically disordered proteins. Annual Review of Biophysics 45, 207231.CrossRefGoogle ScholarPubMed
Schuler, B, Borgia, A, Borgia, MB, Heidarsson, PO, Holmstrom, ED, Nettels, D and Sottini, A (2020) Binding without folding the biomolecular function of disordered polyelectrolyte complexes. Current Opinion in Structural Biology 60, 6676.CrossRefGoogle ScholarPubMed
Sequeira, V, Witjas-Paalberends, ER, Kuster, DWD and van der Velden, J (2014) Cardiac myosin-binding protein C: hypertrophic cardiomyopathy mutations and structure-function relationships. Pflugers Archiv: European Journal of Physiology 466, 201206.CrossRefGoogle ScholarPubMed
Seto, JT, Quinlan, KG, Lek, M, Zheng, XF, Garton, F, MacArthur, DG, Hogarth, MW, Houweling, PJ, Gregorevic, P, Turner, N, Cooney, GJ, Yang, N and North, KN (2013) ACTN3 genotype influences muscle performance through the regulation of calcineurin signaling. The Journal of Clinical Investigation 123, 42554263.CrossRefGoogle ScholarPubMed
Sewanan, LR, Park, J, Rynkiewicz, MJ, Racca, AW, Papoutsidakis, N, Schwan, J, Jacoby, DL, Moore, JR, Lehman, W, Qyang, Y and Campbell, SG (2021) Loss of crossbridge inhibition drives pathological cardiac hypertrophy in patients harboring the TPM1 E192K mutation. Journal of General Physiology 153, e202012640.CrossRefGoogle ScholarPubMed
Shabane, PS, Izadi, S and Onufriev, AV (2019) General purpose water model can improve atomistic simulations of intrinsically disordered proteins. Journal of Chemical Theory and Computation 15, 26202634.CrossRefGoogle ScholarPubMed
Shafaattalab, S, Li, AY, Lin, E, Stevens, CM, Dewar, LJ, Lynn, FC, Sanatani, S, Laksman, Z, Morin, RD, Petegem, Fvan, Hove-Madsen, L, Tieleman, DP, Davis, JP and Tibbits, GF (2019) In vitro analyses of suspected arrhythmogenic thin filament variants as a cause of sudden cardiac death in infants. Proceedings of the National Academy of Sciences 116, 69696974.CrossRefGoogle ScholarPubMed
Sharifi, H, Mann, CK, Rockward, AL, Mehri, M, Mojumder, J, Lee, L-C, Campbell, KS and Wenk, JF (2021) Multiscale simulations of left ventricular growth and remodeling. Biophysical Reviews 13, 729746.CrossRefGoogle ScholarPubMed
Shea, JE, Best, RB and Mittal, J (2021) Physics-based computational and theoretical approaches to intrinsically disordered proteins. Current Opinion in Structural Biology 67, 219225.CrossRefGoogle ScholarPubMed
Sheikh, F, Raskin, A, Chu, P-H, Lange, S, Domenighetti, AA, Zheng, M, Liang, X, Zhang, T, Yajima, T, Gu, Y, Dalton, ND, Mahata, SK, Dorn 2nd GW, Brown, JH, Peterson, KL, Omens, JH, McCulloch, AD and Chen, J (2008) An FHL1-containing complex within the cardiomyocyte sarcomere mediates hypertrophic biomechanical stress responses in mice. Journal of Clinical Investigation 118, 38703880.CrossRefGoogle ScholarPubMed
Shoemaker, BA, Portman, JJ and Wolynes, PG (2000) Speeding molecular recognition by using the folding funnel: the fly-casting mechanism. Proceedings of the National Academy of Sciences 97, 88688873.CrossRefGoogle ScholarPubMed
Shrestha, UR, Juneja, P, Zhang, Q, Gurumoorthy, V, Borreguero, JM, Urban, V, Cheng, X, Pingali, SV, Smith, JC, O’Neill, HM and Petridis, L (2019) Generation of the configurational ensemble of an intrinsically disordered protein from unbiased molecular dynamics simulation. Proceedings of the National Academy of Sciences 116, 2044620452.CrossRefGoogle ScholarPubMed
Sibille, N and Bernado, P (2012) Structural characterization of intrinsically disordered proteins by the combined use of NMR and SAXS. Biochemical Society Transactions 40, 955962.CrossRefGoogle ScholarPubMed
Siddiqui, JK, Tikunova, SB, Walton, SD, Liu, B, Meyer, M, de Tombe, PP, Neilson, N, Kekenes-Huskey, PM, Salhi, HE, Janssen, PM, Biesiadecki, BJ and Davis, JP (2016) Myofilament calcium sensitivity: consequences of the effective concentration of troponin I. Frontiers in Physiology 7, 632.CrossRefGoogle ScholarPubMed
Sielaff, H, Dienerowitz, F and Dienerowitz, M (2022) Single-molecule FRET combined with electrokinetic trapping reveals real-time enzyme kinetics of individual F-ATP synthases. Nanoscale 14, 23272336.CrossRefGoogle ScholarPubMed
Singh, A and Hitchcock-DeGregori, SE (2003) Local destabilization of the tropomyosin coiled coil gives the molecular flexibility required for actin binding. Biochemistry 42, 1411414121.CrossRefGoogle ScholarPubMed
Singh, RR, McNamara, JW and Sadayappan, S (2021) Mutations in myosin S2 alter cardiac myosin-binding protein-C interaction in hypertrophic cardiomyopathy in a phosphorylation-dependent manner. Journal of Biological Chemistry 297, 100836.CrossRefGoogle Scholar
Solis, C and Russell, B (2019) CapZ integrates several signaling pathways in response to mechanical stiffness. Journal of General Physiology 151, 660669.CrossRefGoogle ScholarPubMed
Sols, C and Solaro, RJ (2021) Novel insights into sarcomere regulatory systems control of cardiac thin filament activation. Journal of General Physiology 153, e202012777.CrossRefGoogle Scholar
Song, D, Luo, R and Chen, H-F (2017) The IDP-specific force field ff14IDPSFF improves the conformer sampling of intrinsically disordered proteins. Journal of Chemical Information and Modeling 57, 11661178.CrossRefGoogle ScholarPubMed
Soranno, A, Koenig, I, Borgia, MB, Hofmann, H, Zosel, F, Nettels, D and Schuler, B (2014) Single-molecule spectroscopy reveals polymer effects of disordered proteins in crowded environments. Proceedings of the National Academy of Sciences 111, 48744879.CrossRefGoogle ScholarPubMed
Sormanni, P, Piovesan, D, Heller, GT, Bonomi, M, Kukic, P, Camilloni, C, Fuxreiter, M, Dosztanyi, Z, Pappu, RV, Babu, MM, Longhi, S, Tompa, P, Dunker, AK, Uversky, VN, Tosatto, SCE and Vendruscolo, M (2017) Simultaneous quantification of protein order and disorder. Nature Chemical Biology 13, 339342.CrossRefGoogle ScholarPubMed
Sponga, A, Arolas, JL, Schwarz, TC, Jeffries, CM, Chamorro, AR, Kostan, J, Ghisleni, A, Drepper, F, Polyansky, A, Ribeiro, EDA, Pedron, M, Zawadzka-Kazimierczuk, A, Mlynek, G, Peterbauer, T, Doto, P, Schreiner, C, Hollerl, E, Mateos, B, Geist, L, Faulkner, G, Kozminski, W, Svergun, DI, Warscheid, B, Zagrovic, B, Gautel, M, Konrat, R and Djinovi-Carugo, K (2021) Order from disorder in the sarcomere: FATZ forms a fuzzy but tight complex and phase-separated condensates with α-actinin. Science Advances 7, eabg7653.CrossRefGoogle ScholarPubMed
Spudich, JA, Finer, J, Simmons, B, Ruppel, K, Patterson, B and Uyeda, T (1995) Myosin structure and function. Cold Spring Harbor Symposia on Quantitative Biology 60, 783791.CrossRefGoogle ScholarPubMed
Srivastava, J and Barber, D (2008) Actin co-sedimentation assay; for the analysis of protein binding to F-actin. Journal of Visualized Experiments 13, 690.Google Scholar
Stachowski-Doll, MJ, Papadaki, M, Martin, TG, Ma, W, Gong, HM, Shao, S, Shen, S, Muntu, NA, Kumar, M, Perez, E, Martin, JL, Moravec, CS, Sadayappan, S, Campbell, SG, Irving, T and Kirk, JA (2022) GSK-b2; localizes to the cardiac Z-disc to maintain length dependent activation. Circulation Research 130, 871886.CrossRefGoogle Scholar
Stelzl, LS, Pietrek, LM, Holla, A, Oroz, J, Sikora, M, Köfinger, J, Schuler, B, Zweckstetter, M and Hummer, G (2022) Global structure of the intrinsically disordered protein tau emerges from its local structure. JACS Au 2, 673686.CrossRefGoogle ScholarPubMed
Streng, AS, de Boer, D, van der Velden, J, van Dieijen-Visser, MP and Wodzig, WKWH (2013) Posttranslational modifications of cardiac troponin T: an overview. Journal of Molecular and Cellular Cardiology 63, 4756.CrossRefGoogle ScholarPubMed
Strodel, B (2021) Energy landscapes of protein aggregation and conformation switching in intrinsically disordered proteins. Journal of Molecular Biology 433, 167182.CrossRefGoogle ScholarPubMed
Sucharski, HC, Dudley, EK, Keith, CBR, El Refaey, M, Koenig, SN and Mohler, PJ (2020) Mechanisms and alterations of cardiac ion channels leading to disease: role of ankyrin-B in cardiac function. Biomolecules 10, 211.CrossRefGoogle ScholarPubMed
Sugase, K, Dyson, HJ and Wright, PE (2007) Mechanism of coupled folding and binding of an intrinsically disordered protein. Nature 447, 10211025.CrossRefGoogle ScholarPubMed
Sun, B and Kekenes-Huskey, PM (2020) Molecular basis of S100A1 activation and target regulation within physiological cytosolic Ca2+ levels. Frontiers in Molecular Biosciences 7, 77.CrossRefGoogle Scholar
Sun, B and Kekenes-Huskey, PM (2021) Assessing the role of calmodulin's linker flexibility in target binding. International Journal of Molecular Sciences 22, 4990.CrossRefGoogle ScholarPubMed
Sun, B and Kekenes-Huskey, PM (2022) Calmodulin's interdomain linker is optimized for dynamics signal transmission and calcium binding. Journal of Chemical Information and Modeling 62, 42104221.CrossRefGoogle ScholarPubMed
Sun, B, Blood, R, Atalay, S, Colli, D, Rankin, SE, Barbara, LK and Kekenes-Huskey, PM (2020) Computational Materials, Chemistry, and Biochemistry: From Bold Initiatives to the Last Mile. Cham, Switzerland: Springer, pp. 521558.Google Scholar
Sun, B, Cook, EC, Creamer, TP and Kekenes-Huskey, PM (2018) Electrostatic control of calcineurin's intrinsically-disordered regulatory domain binding to calmodulin. Biochimica et Biophysica Acta – General Subjects 1862, 26512659.CrossRefGoogle ScholarPubMed
Szabo, A, Schulten, K and Schulten, Z (1980) First passage time approach to diffusion controlled reactions. The Journal of Chemical Physics 72, 43504357.CrossRefGoogle Scholar
Szabo, A, Shoup, D, Northrup, SH and McCammon, JA (1982) Stochastically gated diffusioninfluenced reactions. The Journal of Chemical Physics 77, 44844493.CrossRefGoogle Scholar
Szklarczyk, D, Gable, AL, Lyon, D, Junge, A, Wyder, S, Huerta-Cepas, J, Simonovic, M, Doncheva, NT, Morris, JH, Bork, P, Jensen, LJ and Mering, CV (2019) STRING v11: protein-protein association networks with increased coverage, supporting functional discovery in genome-wide experimental datasets. Nucleic Acids Research 47, D607D613.CrossRefGoogle ScholarPubMed
Takeda, S, Yamashita, A, Maeda, K and Maeda, Y (2003) Structure of the core domain of human cardiac troponin in the Ca2+-saturated form. Nature 424, 3541.CrossRefGoogle Scholar
Tande, BM, Wagner, NJ, Mackay, ME, Hawker, CJ and Jeong, M (2001) Viscosimetric, hydrodynamic, and conformational properties of dendrimers and dendrons. Macromolecules 34, 85808585.CrossRefGoogle Scholar
Tanner, BCW, Previs, MJ, Wang, Y, Robbins, J and Palmer, BM (2021) Cardiac myosin binding protein-C phosphorylation accelerates β-cardiac myosin detachment rate in mouse myocardium. American Journal of Physiology Heart and Circulatory Physiology 320, H1822--H1835.CrossRefGoogle ScholarPubMed
Tarnovskaya, S, Kiselev, A, Kostareva, A and Frishman, D (2017) Structural consequences of mutations associated with idiopathic restrictive cardiomyopathy. Amino Acids 49, 18151829.CrossRefGoogle ScholarPubMed
Thangaraju, K, Király, R, Mótyán, JA, Ambrus, VA, Fuxreiter, M and Fésüs, L (2017) Computational analyses of the effect of novel amino acid clusters of human transglutaminase 2 on its structure and function. Amino Acids 49, 605614.CrossRefGoogle ScholarPubMed
Theis, JL, Hu, JJ, Sundsbak, RS, Evans, JM, Bamlet, WR, Qureshi, MY, O’Leary, PW and Olson, TM (2021) Genetic association between hypoplastic left heart syndrome and cardiomyopathies. Circulation: Genomic and Precision Medicine 14, e003126.Google ScholarPubMed
Thomasen, FE, Pesce, F, Roesgaard, MA, Tesei, G and Lindorff-Larsen, K (2022) Improving Martini 3 for disordered and multidomain proteins. Journal of Chemical Theory and Computation 18, 20332041.CrossRefGoogle ScholarPubMed
Timmermann, V, Edwards, AG, Wall, ST, Sundnes, J and McCulloch, AD (2019) Arrhythmogenic current generation by myofilament-triggered Ca2+ release and sarcomere heterogeneity. Biophysical Journal 117, 24712485.CrossRefGoogle ScholarPubMed
Toal, SE, Verbaro, DJ and Schweitzer-Stenner, R (2014) Role of enthalpy–entropy compensation interactions in determining the conformational propensities of amino acid residues in unfolded peptides. The Journal of Physical Chemistry B 118, 13091318.CrossRefGoogle ScholarPubMed
Toepfer, CN, Garfinkel, AC, Venturini, G, Wakimoto, H, Repetti, G, Alamo, L, Sharma, A, Agarwal, R, Ewoldt, JF, Cloonan, P, Letendre, J, Lun, M, Olivotto, I, Colan, S, Ashley, E, Jacoby, D, Michels, M, Redwood, CS, Watkins, HC, Day, SM, Staples, JF, Padrón, R, Chopra, A, Ho, CY, Chen, CS, Pereira, AC, Seidman, JG and Seidman, CE (2020) Myosin sequestration regulates sarcomere function, cardiomyocyte energetics, and metabolism, informing the pathogenesis of hypertrophic cardiomyopathy. Circulation 141, 828842.CrossRefGoogle ScholarPubMed
Tolkatchev, D, Kuruba, B, Smith, GE Jr, Swain, KD, Smith, KA, Moroz, N, Williams, TJ and Kostyukova, AS (2021) Structural insights into the tropomodulin assembly at the pointed ends of actin filaments. Protein Science 30, 423437.CrossRefGoogle ScholarPubMed
Tolman, JR and Ruan, K (2006) NMR residual dipolar couplings as probes of biomolecular dynamics. Chemical Reviews 106, 17201736.CrossRefGoogle ScholarPubMed
Tong, CW, Gaffin, RD, Zawieja, DC and Muthuchamy, M (2004) Roles of phosphorylation of myosin binding protein-C and troponin I in mouse cardiac muscle twitch dynamics. The Journal of Physiology 558, 927941.CrossRefGoogle ScholarPubMed
Towns, J, Cockerill, T, Dahan, M, Foster, I, Gaither, K, Grimshaw, A, Hazlewood, V, Lathrop, S, Lifka, D, Peterson, GD, Roskies, R, Scott, JR and Wilkins-Diehr, N (2014) XSEDE: accelerating scientific discovery. Computing in Science & Engineering 16, 6274.CrossRefGoogle Scholar
Trayanova, N and Rice, J (2011) Cardiac electromechanical models: from cell to organ. Frontiers in Physiology 2, 43.CrossRefGoogle ScholarPubMed
Tsukada, T, Pappas, CT, Moroz, N, Antin, PB, Kostyukova, AS and Gregorio, CC (2010) Leiomodin-2 is an antagonist of tropomodulin-1 at the pointed end of the thin filaments in cardiac muscle. Journal of Cell Science 123(Pt 18), 31363145.CrossRefGoogle ScholarPubMed
Tucholski, T, Cai, W, Gregorich, ZR, Bayne, EF, Mitchell, SD, McIlwain, SJ, Lange, WJde, Wrobbel, M, Karp, H, Hite, Z, Vikhorev, PG, Marston, SB, Lal, S, Li, A, Remedios, Cdos, Kohmoto, T, Hermsen, J, Ralphe, JC, Kamp, TJ, Moss, RL and Ge, Y (2020) Distinct hypertrophic cardiomyopathy genotypes result in convergent sarcomeric proteoform profiles revealed by top-down proteomics. Proceedings of the National Academy of Sciences 117, 2469124700.CrossRefGoogle ScholarPubMed
Turoverov, KK, Kuznetsova, IM and Uversky, VN (2010) The protein kingdom extended: ordered and intrinsically disordered proteins, their folding, supramolecular complex formation, and aggregation. Progress in Biophysics and Molecular Biology 102, 7384.CrossRefGoogle ScholarPubMed
Uversky, VN (2009) Intrinsically disordered proteins and their environment: effects of strong denaturants, temperature, pH, counter ions, membranes, binding partners, osmolytes, and macromolecular crowding. The Protein Journal 28, 305325.CrossRefGoogle ScholarPubMed
Uversky, VN (2014) Wrecked regulation of intrinsically disordered proteins in diseases: pathogenicity of deregulated regulators. Frontiers in Molecular Biosciences 1, 6.CrossRefGoogle ScholarPubMed
Uversky, VN (2020) Intrinsically disordered proteins. In Renaud, J-P (ed.), Structural Biology in Drug Discovery. John Wiley & Sons, Ltd, pp. 587612.CrossRefGoogle Scholar
Uversky, VN, Oldfield, CJ and Dunker, AK (2005) Showing your ID: intrinsic disorder as an ID for recognition, regulation and cell signaling. Journal of Molecular Recognition 18, 343384.CrossRefGoogle ScholarPubMed
Uversky, VN, Shah, SP, Gritsyna, Y, Hitchcock-DeGregori, SE and Kostyukova, AS (2011) Systematic analysis of tropomodulin/tropomyosin interactions uncovers fine-tuned binding specificity of intrinsically disordered proteins. Journal of Molecular Recognition 24, 647655.CrossRefGoogle ScholarPubMed
Vacic, V and Iakoucheva, LM (2012) Disease mutations in disordered regions–exception to the rule?. Molecular BioSystems 8, 2732.CrossRefGoogle ScholarPubMed
van der Velden, J and Stienen, GJ (2019) Cardiac disorders and pathophysiology of sarcomeric proteins. Physiological Reviews 99, 381426.CrossRefGoogle ScholarPubMed
Van Valen, D, Haataja, M and Phillips, R (2009) Biochemistry on a leash: the roles of tether length and geometry in signal integration proteins. Biophysical Journal 96, 12751292.CrossRefGoogle ScholarPubMed
Velázquez-Campoy, A, Ohtaka, H, Nezami, A, Muzammil, S and Freire, E (2004) Isothermal titration calorimetry. Current Protocols in Cell Biology 23, 17.8.117.8.24.CrossRefGoogle Scholar
Vijaykumar, A, Bolhuis, PG and Ten Wolde, PR (2016) The intrinsic rate constants in diffusion-influenced reactions. Faraday Discussions 195, 421441.CrossRefGoogle ScholarPubMed
Vitalis, A and Pappu, RV (2009) ABSINTH: a new continuum solvation model for simulations of polypeptides in aqueous solutions. Journal of Computational Chemistry 30, 673699.CrossRefGoogle ScholarPubMed
Vite, A, Li, J and Radice, GL (2015) New functions for alpha-catenins in health and disease: from cancer to heart regeneration. Cell and Tissue Research 360, 773783.CrossRefGoogle ScholarPubMed
von der Ecken, J, Heissler, SM, Pathan-Chhatbar, S, Manstein, DJ and Raunser, S (2016) Cryo-EM structure of a human cytoplasmic actomyosin complex at near-atomic resolution. Nature 534, 724728.CrossRefGoogle ScholarPubMed
Votapka, LW and Amaro, RE (2015) Multiscale estimation of binding kinetics using Brownian dynamics, molecular dynamics and milestoning. PLOS Computational Biology 11, 124.CrossRefGoogle ScholarPubMed
Votapka, LW, Jagger, BR, Heyneman, AL and Amaro, RE (2017) SEEKR: simulation enabled estimation of kinetic rates, a computational tool to estimate molecular kinetics and its application to trypsin–benzamidine binding. The Journal of Physical Chemistry B 121, 35973606.CrossRefGoogle Scholar
Votapka, LW, Stokely, AM, Ojha, AA and Amaro, RE (2022) SEEKR2: versatile multiscale milestoning utilizing the OpenMM molecular dynamics engine. Journal of Chemical Information and Modeling 62, 32533262.CrossRefGoogle ScholarPubMed
Vu, LD, Gevaert, K and De Smet, I (2018) Protein language: post-translational modifications talking to each other. Trends in Plant Science 23, 10681080.CrossRefGoogle ScholarPubMed
Wagner, W, Fodor, E, Ginsburg, A and Hammer, JA (2006) The binding of DYNLL2 to myosin Va requires alternatively spliced exon B and stabilizes a portion of the myosin's coiled-coil domain. Biochemistry 45, 1156411577.CrossRefGoogle ScholarPubMed
Wang, W (2021) Recent advances in atomic molecular dynamics simulation of intrinsically disordered proteins. Physical Chemistry Chemical Physics 23, 777784.CrossRefGoogle ScholarPubMed
Wang, Y, Chu, X, Longhi, S, Roche, P, Han, W, Wang, E and Wang, J (2013) Multiscaled exploration of coupled folding and binding of an intrinsically disordered molecular recognition element in measles èirus nucleoprotein. Proceedings of the National Academy of Sciences 110, E3743E3752.Google Scholar
Wang, C, Wei, Z, Chen, K, Ye, F, Yu, C, Bennett, V and Zhang, M (2014) Structural basis of diverse membrane target recognitions by ankyrins. eLife 3, e04353.CrossRefGoogle ScholarPubMed
Wang, J, Peng, C, Yu, Y, Chen, Z, Xu, Z, Cai, T, Shao, Q, Shi, J and Zhu, W (2020) Exploring conformational change of adenylate kinase by replica exchange molecular dynamic simulation. Biophysical Journal 118, 10091018.CrossRefGoogle ScholarPubMed
Wang, Z, Grange, M, Wagner, T, Kho, AL, Gautel, M and Raunser, S (2021) The molecular basis for sarcomere organization in vertebrate skeletal muscle. Cell 184, 21352150.e13.CrossRefGoogle ScholarPubMed
Weiner, SJ, Kollman, PA, Case, DA, Singh, UC, Ghio, C, Alagona, G, Profeta, S, and Weiner, P (1984) A new force field for molecular mechanical simulation of nucleic acids and proteins. Journal of the American Chemical Society 106, 765784.CrossRefGoogle Scholar
Wei, T-C, Lin, H-Y, Lu, C-C, Chen, C-M and You, L-R (2011) Expression of Crip2, a LIM-domain-only protein, in the mouse cardiovascular system under physiological and pathological conditions. Gene Expression Patterns 11, 384394.CrossRefGoogle ScholarPubMed
Whitley, JA, Ex-Willey, AM, Marzolf, DR, Ackermann, MA, Tongen, AL, Kokhan, O and Wright, NT (2019) Obscurin is a semi-flexible molecule in solution. Protein Science 28, 717726.CrossRefGoogle ScholarPubMed
Wicky, BIM, Shammas, SL and Clarke, J (2017) Affinity of IDPs to their targets is modulated by ion-specific changes in kinetics and residual structure. Proceedings of the National Academy of Sciences 114, 98829887.CrossRefGoogle ScholarPubMed
Willis, MS, Schisler, JC, Portbury, AL and Patterson, C (2008) Build it up-tear it down: protein quality control in the cardiac sarcomere. Cardiovascular Research 81, 439448.CrossRefGoogle ScholarPubMed
Willott, RH, Gomes, AV, Chang, AN, Parvatiyar, MS, Pinto, JR and Potter, JD (2010) Mutations in troponin that cause HCM, DCM AND RCM: what can we learn about thin filament function? Journal of Molecular and Cellular Cardiology 48, 882892.CrossRefGoogle ScholarPubMed
Winkelmann, JC, Chang, JG, Tse, WT, Scarpa, AL, Marchesi, VT and Forget, BG (1990) Full-length sequence of the cDNA for human erythroid beta-spectrin. Journal of Biological Chemistry 265, 1182711832.CrossRefGoogle ScholarPubMed
Wohl, S, Jakubowski, M and Zheng, W (2021) Salt-dependent conformational changes of intrinsically disordered proteins. Journal of Physical Chemistry Letters 12, 66846691.CrossRefGoogle ScholarPubMed
Wright, PE and Dyson, HJ (2015) Intrinsically disordered proteins in cellular signalling and regulation. Nature Reviews Molecular Cell Biology 16, 1829.CrossRefGoogle ScholarPubMed
Wu, M-C, Forbes, JG and Wang, K (2016) Disorder profile of nebulin encodes a vernierlike position sensor for the sliding thin and thick filaments of the skeletal muscle sarcomere. Physical Review E 93, 062406.CrossRefGoogle ScholarPubMed
Wuttke, R, Hofmann, H, Nettels, D, Borgia, MB, Mittal, J, Best, RB and Schuler, B (2014) Temperature-dependent solvation modulates the dimensions of disordered proteins. Proceedings of the National Academy of Sciences 111, 52135218.CrossRefGoogle ScholarPubMed
Wu, K-P, Weinstock, DS, Narayanan, C, Levy, RM and Baum, J (2009) Structural reorganization of α-synuclein at low pH observed by NMR and REMD simulations. Journal of Molecular Biology 391, 784796.CrossRefGoogle Scholar
Xue, B, Dunbrack, RL, Williams, RW, Dunker, AK and Uversky, VN (2010) PONDR-FIT: a meta-predictor of intrinsically disordered amino acids. Biochimica et Biophysica Acta 1804, 9961010.CrossRefGoogle ScholarPubMed
Xu, Y, Zheng, Y, Fan, J-S and Yang, D (2006) A new strategy for structure determination of large proteins in solution without deuteration. Nature Methods 3, 931937.CrossRefGoogle ScholarPubMed
Xu, Q, Dewey, S, Nguyen, S and Gomes, AV (2010) Malignant and benign mutations in familial cardiomyopathies: insights into mutations linked to complex cardiovascular phenotypes. Journal of Molecular and Cellular Cardiology 48, 899909.CrossRefGoogle ScholarPubMed
Yamada, Y, Namba, K and Fujii, T (2020) Cardiac muscle thin filament structures reveal calcium regulatory mechanism. Nature Communications 11, 153.CrossRefGoogle ScholarPubMed
Yamasaki, R, Berri, M, Wu, Y, Trombitas, K, McNabb, M, Kellermayer, MS, Witt, C, Labeit, D, Labeit, S, Greaser, M and Granzier, H (2001) Titin-actin interaction in mouse myocardium: passive tension modulation and its regulation by calcium/S100A1. Biophysical Journal 81, 22972313.CrossRefGoogle ScholarPubMed
Yang, Y and Igumenova, TI (2013) The C-terminal V5 domain of protein kinase Ca is intrinsically disordered, with propensity to associate with a membrane mimetic. PLoS ONE 8, e65699.CrossRefGoogle Scholar
Yang, H, Xiao, X, Zhao, X and Wu, Y (2015) Intrinsic fluorescence spectra of tryptophan, tyrosine and phenyloalanine. Proceedings of the 5th International Conference on Advanced Design and Manufacturing Engineering. Paris, France: Atlantis Press.CrossRefGoogle Scholar
Yang, J, Gao, M, Xiong, J, Su, Z and Huang, Y (2019) Features of molecular recognition of intrinsically disordered proteins via coupled folding and binding. Protein Science 28, 19521965.CrossRefGoogle ScholarPubMed
Yar, S, Monasky, MM and Solaro, RJ (2014) Maladaptive modifications in myofilament proteins and triggers in the progression to heart failure and sudden death. Pflugers Archiv: European Journal of Physiology 466, 11891197.CrossRefGoogle ScholarPubMed
Yar, S, Chowdhury, SA, Davis, RT, Kobayashi, M, Monasky, MM, Rajan, S, Wolska, BM, Gaponenko, V, Kobayashi, T, Wieczorek, DF and Solaro, RJ (2013) Conserved Asp-137 is important for both structure and regulatory functions of cardiac α-tropomyosin (α-TM) in a novel transgenic mouse model expressing α-TM-D137L*. Journal of Biological Chemistry 288, 1623516246.CrossRefGoogle Scholar
Yeates, TO, Agdanowski, MP and Liu, Y (2020) Development of imaging scaffolds for cryo-electron microscopy. Current Opinion in Structural Biology 60, 142149.CrossRefGoogle ScholarPubMed
Young, P, Ehler, E and Gautel, M (2001) Obscurin, a giant sarcomeric Rho guanine nucleotide exchange factor protein involved in sarcomere assembly. Journal of Cell Biology 154, 123136.CrossRefGoogle ScholarPubMed
Yu, W, Huang, M-M, Zhang, G-H, Wang, W, Chen, C-J and Cheng, J-D (2021) Whole-exome sequencing reveals MYH7 p.R671C mutation in three different phenotypes of familial hypertrophic cardiomyopathy. Experimental and Therapeutic Medicine 22, 1002.CrossRefGoogle ScholarPubMed
Zabrouskov, V, Ge, Y, Schwartz, J and Walker, JW (2008) Unraveling molecular complexity of phosphorylated human cardiac troponin I by top down electron capture dissociation/electron transfer dissociation mass spectrometry. Molecular & Cellular Proteomics 7, 18381849.CrossRefGoogle ScholarPubMed
Zamora, JE, Papadaki, M, Messer, AE, Marston, SB and Gould, IR (2016) Troponin structure: its modulation by Ca2+ and phosphorylation studied by molecular dynamics simulations. Physical Chemistry Chemical Physics 18, 2069120707.CrossRefGoogle ScholarPubMed
Zhang, Y and Sagui, C (2015) Secondary structure assignment for conformationally irregular peptides: comparison between DSSP, STRIDE and KAKSI. Journal of Molecular Graphics and Modelling 55, 7284.CrossRefGoogle ScholarPubMed
Zhang, W, Wu, C and Duan, Y (2005) Convergence of replica exchange molecular dynamics. The Journal of Chemical Physics 123, 154105.CrossRefGoogle ScholarPubMed
Zhang, H, Rajasekaran, NS, Orosz, A, Xiao, X, Rechsteiner, M and Benjamin, IJ (2010) Selective degradation of aggregate-prone CryAB mutants by HSPB1 is mediated by ubiquitin-proteasome pathways. Journal of Molecular and Cellular Cardiology 49, 918930.CrossRefGoogle ScholarPubMed
Zhao, YG and Zhang, H (2020) Phase separation in membrane biology: the interplay between membrane-bound organelles and membraneless condensates. Developmental Cell 55, 3044.CrossRefGoogle ScholarPubMed
Zheng, W and Best, RB (2018) An extended Guinier analysis for intrinsically disordered proteins. Journal of Molecular Biology 430, 25402553.CrossRefGoogle ScholarPubMed
Zheng, X, Bi, C, Li, Z, Podariu, M and Hage, DS (2015) Analytical methods for kinetic studies of biological interactions: a review. Journal of Pharmaceutical and Biomedical Analysis 113, 163180.CrossRefGoogle ScholarPubMed
Zhou, HX (2001) The affinity-enhancing roles of flexible linkers in two-domain DNA-binding proteins. Biochemistry 40, 1506915073.CrossRefGoogle ScholarPubMed
Zhou, H-X (2008) Effect of mixed macromolecular crowding agents on protein folding. Proteins: Structure, Function, and Bioinformatics 72, 11091113.CrossRefGoogle ScholarPubMed
Zhou, T, Fleming, JR, Lange, S, Hessel, AL, Bogomolovas, J, Stronczek, C, Grundei, D, Ghassemian, M, Biju, A, Borgeson, E, Bullard, B, Linke, WA, Chen, J, Kovermann, M and Mayans, O (2021 a) Molecular characterisation of titin N2A and its binding of CARP reveals a titin/actin cross-linking mechanism. Journal of Molecular Biology 433, 166901.CrossRefGoogle ScholarPubMed
Zhou, X, Lin, Y, Kato, M, Mori, E, Liszczak, G, Sutherland, L, Sysoev, VO, Murray, DT, Tycko, R and McKnight, SL (2021 b) Transiently structured head domains control intermediate filament assembly. Proceedings of the National Academy of Sciences 118, e2022121118.CrossRefGoogle ScholarPubMed
Zimm, BH, Stockmayer, WH and Fixman, M (1953) Excluded volume in polymer chains. The Journal of Chemical Physics 21, 17161723.CrossRefGoogle Scholar
Zwier, MC, Pratt, AJ, Adelman, JL, Kaus, JW, Zuckerman, DM and Chong, LT (2016) Efficient atomistic simulation of pathways and calculation of rate constants for a protein-peptide binding process: application to the MDM2 protein and an intrinsically disordered p53 peptide. The Journal of Physical Chemistry Letters 7, 34403445.CrossRefGoogle Scholar
Figure 0

Fig. 1. (a) Schematic illustration of the sarcomere (drawn with BioRender). (b) In this review, we focus on the cardiac proteins proposed in Kooij et al. (2014) with some additional noteworthy examples. For proteins with multiple isoforms, only isoforms with spectra counts (SC) >10 were selected. Proteins with IDR(s) are indicated by red *. Double ** indicates that the IDR(s) has been experimentally confirmed for the gene, while a single * indicates that the confirmation was based on a related isoform or via bioinformatic predictions (see Table 1 for details).

Figure 1

Fig. 2. Core proteins of the thin and thick filament, based on the schematic from Harris et al. (2011). The thin filament structure PDB 6KN8 was constructed from a cryo-EM study (Yamada et al., 2020). PDB 5TBY was generated from homology modeling. The MyBPC3 structure was predicted by AlphaFold and was downloaded from the UniprotKB database. A 2 nm resolution model of tarantula thick filament was built by fitting atomistic component structures to EM density map (PDB 3DTP (Alamo et al., 2008)).

Figure 2

Fig. 3. (a) The IDP-phase diagram developed by the Pappu lab (Holehouse et al., 2017), which groups proteins by characteristic disorder including molten, extended, or compact (Uversky, 2020) classes have also been proposed based on their charge patterns (R1–R5) (Holehouse et al., 2017): R1 corresponds to weak polyampholytes and resembles pre-molten globules. R3 signifies strong polyampholytes with a comparable amount of positively and negatively charged residues and is described as hairpins/coils/chimeras. R2 is the boundary between R1 and R3 where coils and pre-molten globules coexist. R4 and R5 are strong polyampholytes like R3, but with dominant negative and positive residues, respectively. IDPs in R4 and R5 are swollen coils. (be) PONDR-VLXT predicted disordered regions in cardiac myofilament proteins. These proteins are categorized into thin/thick filament(s), Z-disk, and miscellaneous. The IDP region is colored red and interlaced with folded regions. The blue line depicts the first and last amino acid and the number is increasing counterclockwise. The numbers in the parentheses present the percentage of predicted IDR residues, ‘pathogenic or likely pathogenic’ mutations, and phosphorylation sites located in the predicted IDP regions, respectively. The structural state estimation of predicted >5 residue IDR regions before (black dots) and after phosphorylation (magenta dots, if PTM site exists in the IDP region) in the IDP-phase diagram were also shown.

Figure 3

Fig. 4. Solution NMR structures of the common modular domains that are used as building blocks of Z-disk proteins. These structures show highly dynamic termini and evidence of intrinsic disorder in the Z-disk. The PDB IDs are given in parentheses.

Figure 4

Fig. 5. (a) Major secondary structural elements present in folded proteins like actin (PDB 6KN8 (Yamada et al., 2020)). (b) NMR structures of a MAPID, the MyBPC3 construct consisting of the M- and C2 domains (Michie et al., 2016), are shown as an example to illustrate the IDR ensemble. (c) Scaling of radius of gyration (Rg) versus chain length for proteins at different states (Lazar et al., 2021). (d) The IDP phase-diagram for classifying IDPs into five structural states (R1–R5) based on the charge patterns (Holehouse et al., 2017).

Figure 5

Fig. 6. Schematic of transient and covalent influences on IDR structure. Ions can screen intramolecular electrostatics or directly coordinate with charged residues. pH affects the protonation state of ionizable residues (e.g. histidine). Phosphorylation introduces negative charges into the sequence. For crowding, many factors such as size and surface charges of crowders, and volume fraction affect IDR structure (Eq. (7)).

Figure 6

Table 1. Brief summary of reported experimental and computational studies on MAPIDs

Figure 7

Fig. 7. Computational methods commonly used to model IDR structural properties. The IDR propensity of a structure can be predicted from its amino acid sequence by tens of established tools (Liu et al., 2019), and structural states of IDP can be either explicitly modeled (Ozenne et al., 2012) or characterized by an implicit phase diagram (Holehouse et al., 2017). Polymer models including the simplistic freely jointed monomer model and advanced models that account for intramonomer interactions can be used to characterize IDR ensembles (Milstein and Meiners, 2013; Schuler et al., 2016). Particle-based simulations are commonly used to predict IDR conformer structures and associated kinetics (Wang, 2021). All atom simulations consider every individual atom in the system to determine detailed descriptions of the potential energy surface (PES) but are computationally intensive, while coarse-grained simulations lump atoms together to increase sampling efficiency at a modest loss of accuracy. Lastly, statistical models frequently use partition functions to obtain thermodynamic descriptions of IDRs (Hadzi et al., 2021).

Figure 8

Fig. 8. (a) Dynamics of an IDP ensemble represented by a Markov state model (MSM). (b) Experimental methods for structure determination and their temporal resolutions. Timescale information of NMR, X-ray, cryo-EM, and SAXS are taken from Ban (2020). FRET time resolution spans ns to seconds (Okamoto and Sako, 2017). Size information: EM is appropriate for proteins >100 kD (Yeates et al., 2020). NMR is most suitable for <30 kD proteins (Xu et al., 2006). X-ray can solve structures up to 4000 kD (PDB website statistics). FRET is used for medium-sized proteins but has been reported for up to a 540 kD protein (Sielaff et al., 2022).

Figure 9

Fig. 9. (a) Binding of IDP to its partner often goes through a coupled-folding-and-binding process (Sugase et al., 2007), in which both the intramolecular conversion kinetics (section ‘Computational methods for predicting intramolecular dynamics of MAPIDs’) of the IDP and its intermolecular association kinetics (section ‘Computational methods for predicting the MAPID co-assembly’) are important. (b) Theoretical frameworks for describing IDP intermolecular kinetics. The Smoluchowski equation is an approximation for a diffusion-limited association rate (Kim et al., 2018). The Van Valen et al. model (Van Valen et al., 2009) (biochemistry on a leash) combines IDP-enhanced effective concentrations and competitive binding to describe IDP/target binding (Eqs. (10) and (56)). The fly-casting model (Shoemakeret al., 2000) explains the kinetic advantage of IDP/target assembly through its fast searching for binding partners. The Brownian dynamics (browndye (Huber and McCammon, 2010)) and the SEEKR (Votapka et al., 2017) programs both use the Northrup Allison McCammon algorithm (Northrup et al., 1984) provide simulation-based estimates of association kinetics and thermodynamics quantities.

Supplementary material: PDF

Sun and Kekenes-Huskey supplementary material

Sun and Kekenes-Huskey supplementary material

Download Sun and Kekenes-Huskey supplementary material(PDF)
PDF 442.9 KB