Next Article in Journal
In Vitro Estimation of Relative Compliance during High-Frequency Oscillatory Ventilation
Next Article in Special Issue
The Challenging Riddle about the Janus-Type Role of Hsp60 and Related Extracellular Vesicles and miRNAs in Carcinogenesis and the Promises of Its Solution
Previous Article in Journal
Experimental Evaluation of a Spectral Index to Characterize Temporal Variations in the Direct Normal Irradiance Spectrum
Previous Article in Special Issue
Functions and Therapeutic Potential of Extracellular Hsp60, Hsp70, and Hsp90 in Neuroinflammatory Disorders
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Neurochaperonopathies: Anomalies of the Chaperone System with Pathogenic Effects in Neurodegenerative and Neuromuscular Disorders

by
Federica Scalia
1,2,
Alessandra Maria Vitale
1,2,
Radha Santonocito
1,
Everly Conway de Macario
3,
Alberto J. L. Macario
2,3 and
Francesco Cappello
1,2,*
1
Department of Biomedicine, Neuroscience and Advanced Diagnostics (BIND), University of Palermo, 90127 Palermo, Italy
2
Euro-Mediterranean Institute of Science and Technology (IEMEST), 90139 Palermo, Italy
3
Department of Microbiology and Immunology, School of Medicine, University of Maryland at Baltimore-Institute of Marine and Environmental Technology (IMET), Baltimore, MD 21202, USA
*
Author to whom correspondence should be addressed.
Appl. Sci. 2021, 11(3), 898; https://doi.org/10.3390/app11030898
Submission received: 28 December 2020 / Revised: 10 January 2021 / Accepted: 12 January 2021 / Published: 20 January 2021

Abstract

:
The chaperone (or chaperoning) system (CS) constitutes molecular chaperones, co-chaperones, and chaperone co-factors, interactors and receptors, and its canonical role is protein quality control. A malfunction of the CS may cause diseases, known as the chaperonopathies. These are caused by qualitatively and/or quantitatively abnormal molecular chaperones. Since the CS is ubiquitous, chaperonopathies are systemic, affecting various tissues and organs, playing an etiologic-pathogenic role in diverse conditions. In this review, we focus on chaperonopathies involved in the pathogenic mechanisms of diseases of the central and peripheral nervous systems: the neurochaperonopathies (NCPs). Genetic NCPs are linked to pathogenic variants of chaperone genes encoding, for example, the small Hsp, Hsp10, Hsp40, Hsp60, and CCT-BBS (chaperonin-containing TCP-1- Bardet–Biedl syndrome) chaperones. Instead, the acquired NCPs are associated with malfunctional chaperones, such as Hsp70, Hsp90, and VCP/p97 with aberrant post-translational modifications. Awareness of the chaperonopathies as the underlying primary or secondary causes of disease will improve diagnosis and patient management and open the possibility of investigating and developing chaperonotherapy, namely treatment with the abnormal chaperone as the main target. Positive chaperonotherapy would apply in chaperonopathies by defect, i.e., chaperone insufficiency, and consist of chaperone replacement or boosting, whereas negative chaperonotherapy would be pertinent when a chaperone actively participates in the initiation and progression of the disease and must be blocked and eliminated.

Graphical Abstract

1. Introduction

A better understanding of neurodegenerative diseases may be achieved by examining them as part of the broad area of protein quality control, since they show signs of protein pathology. Indeed, a common feature in neuropathology is protein misfolding with subsequent formation of protein aggregates and precipitates, and various neurological disorders are proteinopathies or have a proteinopathy component [1,2,3,4]. In these, one or more proteins are pathogenic and prone to aggregate because of a genetic or acquired abnormality. In genetic proteinopathies, the gene encoding the pathologic protein bears a variant which makes its product pathogenic, whereas in the acquired proteinopathies the gene is normal, but the encoded protein may undergo a post-translation modification (PTM). In either case, aggregates formed by some of the abnormal proteins are cytotoxic and cause disease. Through evolution, humans, as well as other organisms, have developed mechanisms to deal with proteinopathies. These mechanisms include protein-degradation and autophagic machineries, for instance the ubiquitin-proteasome system (UPS) [5,6,7] and chaperone-mediated autophagy (CMA) [8,9,10]. In addition, organisms have developed means to ensure that all proteins fold correctly, translocate to the place in which they function, resist the damaging effect of stressors, and regain activity after reversible damage and aggregation, namely all mechanisms that maintain protein homeostasis. Major players in protein-quality control are a variety of molecular chaperones, some of which are called Hsp (heat shock protein) that form groups of phylogenetically related proteins, such as the small-heat shock protein (sHsp), Hsp40(DnaJ), Hsp70(DnaK), Hsp90, and heavy Hsp families. In addition, there is a class of protein chaperones of about 60 kDa, also called chaperonins, of which two groups are distinguished, I (Hsp60, or Cpn60 for chaperonin 60) and II (CCT for chaperonin-containing TCP-1; also called TriC, for TCP-1 ring complex). All these chaperones typically function as multimolecular complexes of various sizes and degrees of complexity [11,12,13,14,15,16,17] and are part of the chaperone system (CS), which also includes co-chaperones, chaperone co-factors, and chaperone interactors and receptors [18,19,20]. The main functional partners of the CS in protein-quality control are the UPS and the CMA machineries. When the CS malfunctions, for example because of an abnormal chaperone, disease may ensue: a chaperonopathy [18,21] (free access updates at http://www.chaperones-pathology.org/). In many instances chaperonopathies have a pathogenic impact on the nervous system and contribute to the mechanisms underpinning neuropathies [20,22,23,24]. In this article, we will briefly discuss examples of neuropathies in which chaperonopathies play an etiologic-pathogenic role. These conditions we will call neurochaperonopathies (NCPs).

2. Types of Chaperonopathies in NCPs

NCPs are a heterogeneous group of diseases with complicated etiological and pathogenic features, among which failure of the CS can contribute to various degrees, depending on the disease and the patient considered. Here, we will briefly discuss some salient aspects of NCPs we considered useful to pathologists and physicians for understanding these conditions and for managing these patients.
The canonical functions of the CS pertain to maintenance of protein homoeostasis together with the CMA and the UPS, while its non-canonical functions are unrelated with the former and mostly co-involve the immune system (IS) [20,21]. A deficiency in a component of the CS, for example a chaperone, may cause disease, a chaperonopathy, which can be genetic or acquired [18,21]. In the former, a genetic variant, for example a mutation, results in the production of a defective, malfunctional chaperone, which may lead to tissue lesions and disease (illustrative examples and pertinent bibliography are provided in Table 1). Instead, in acquired chaperonopathies the gene encoding the affected chaperone is normal, but the gene product, namely the protein chaperone, is not. Chaperone abnormalities, genetic and acquired can also be classified considering the level of the chaperone’s concentration and/or physiological activity by defect, excess, and mistake. In the chaperonopathies by defect, a chaperone deficit is the distinctive feature, which can be qualitative, quantitative or both. In the former, the affected chaperone may be at the right concentration, but its function is impaired because of a structural defect, for example; or the chaperone may be at a concentration below the normal range and it cannot reach the physiological level of functionality; or both deficits may coincide, below-normal concentration and functional ability, a situation often occurring in genetic chaperonopathies. In this regard, it should be emphasized that in neuropathies with abundant protein aggregates and precipitates, the level of chaperones available to deal with the excessive demand from misfolding clients may not be sufficient. This generates a chaperonopathy by defect caused by a disproportionate demand even when chaperone production is normal. The pool of useful chaperone molecules may be even further depleted by their being sequestered in the aggregates and precipitates [21] (free access updates at http://www.chaperones-pathology.org/). The opposite parameters occur in the chaperonopathies by excess, namely the affected chaperone is either quantitatively above the normal range of concentrations or is qualitatively abnormal, displaying abnormal functions (gain of function).
The chaperonopathies by mistake include disorders in which a chaperone, apparently normal according to the results of tests currently available, plays a pathogenic role: it participates in the mechanism of disease as it occurs in various types of cancer and inflammatory and autoimmune conditions. In these instances, the pathogenic chaperone may have been structurally modified post-transcriptionally, for example by a post-translation modification (PTM), and thereby its intrinsic properties and functions changed. This is a field that calls for intensive research to develop means to detect PTMs and measure their impact on the chaperone molecule’s properties and functions in single cells or using minimal amounts of chaperone. A case in point is the chaperone Hsp60, which plays a variety of roles, canonical and non-canonical, physiological, and pathogenic, even though there is only a single Hsp60 gene in the human genome. How this multifunctionality is generated when there is only a single Hsp60 molecular type? One mechanism could be through post-translational modifications, a possibility that deserves further investigation [25].
All the above-mentioned types of chaperonopathies, involving the chaperones listed in Table 1 have been implicated in neurological disorders and some examples of these NCPs are discussed below.

3. Genetic NCPs

3.1. Small Heat Shock Proteins

Diverse neurological and muscular disorders have been associated with mutations in sHsps. The sHsp α-crystallin family includes 10 members and one related protein (HSP16.2/HSPB11) [66] that, in addition to the chaperoning function, participate in cytoskeleton stabilization and possess anti-aggregation and anti-apoptotic activities [67,68,69,70,71].
Most of the mutations found in the sHsps are located in the highly conserved α-crystallin domain, which is an 80–100 amino-acid-long domain responsible for the association/dissociation of sHsp dimers and for the formation and stabilization of large multisubunit homo- and hetero-oligomers [72]. These mutations lead to protein aggregation and cell death. Two hypotheses have been proposed to explain the pathologic findings: the mutated sHsp molecules may acquire an intrinsic toxicity (gain of toxic function), or they may suffer a loss of function with abolition of their protein quality control activity [73]. Two representative examples are the mutations occurring in the HSPB8 and HSPB1 genes that have been associated with the neuromuscular disorders Charcot-Marie-Tooth (CMT) disease and distal hereditary motor neuropathies (dHMN) (Table 1). The term CMT is used to indicate a clinically and genetically heterogeneous group of hereditary motor and sensory neuropathies, characterized by degeneration of peripheral nerves and subdivided in two subgroups: demyelinating (CMT1) and axonal (CMT2) [74]. dHMNs are genetically heterogenous characterized by degeneration of distal lower motor neurons, resulting in muscle weakness and atrophy [75]. However, in many forms of dHMN there may also be minor sensory abnormalities and/or a significant upper-motor-neuron involvement. For this reason, there is often an overlap not only with axonal forms of CMT, i.e., CMT2, but also with juvenile forms of amyotrophic lateral sclerosis, Kennedy’s disease, spinal muscular atrophy, and hereditary spastic paraplegia [76].
Various types of chaperonopathies can occur, as follows: CMT type 2L is caused by mutations in HSPB8 and CMT type 2F is caused by mutations in HSPB1, while dHMN type IIA is caused by mutations in HSPB8 and dHMN type IIB is caused by mutations in HSPB1 [27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44]. Often, a single mutation in HSPB1 or HSPB8 gives rise to one of these conditions. However, in many cases the same mutation was found associated with both diseases, such as the p.(Lys141Asn) substitution in HSPB8 [28,31], and the p.(Ser135Phe) [38,54] and p.(Arg127Trp) substitutions in HSPB1 [36,37,77]. Moreover, there are also examples of mutations found in patients with an unclear phenotype, as shown in Table 1.
For both HSPB8 and HSPB1, the identified mutations in most cases occur in the highly conserved α-crystallin domain, likely destabilizing the proteins structurally and functionally [78]. To elucidate the molecular mechanisms responsible for the association between mutated forms of HSPB8 and HSPB1 and neuromuscular disorders, in vitro studies have been performed. For instance, it was reported that the expression of HSPB8 mutants p.(Lys141Asn) and p.(Lys141Glu) in motor neurons resulted in a reduction in average length and number of neurites per cell, without inducing cell death. In contrast, these abnormalities were very moderate in sensory neurons, and absent in cortical neurons and glial cells [79]. Thus, motor neurons appear to be more sensitive to HSPB8 dysfunction compared to sensory neurons, as indicated by the predominant motor neuron phenotype in dHMN and CMT2L. Moreover, both HSPB8 mutant forms tend to associate with HSPB1 mutant proteins, forming intracellular aggregates [28].
Studies in vitro with HSPB1 mutants demonstrated that some of the substitutions affecting the α-crystallin domain reduced motor neuron viability, impaired neurofilament assembly, destabilized microtubules, and disrupted the axonal transport of specific cellular cargoes and of mitochondria [36,80,81,82,83,84], events that can be responsible for the premature axonal degeneration, typical of both CMT and dHMN [85]. In addition, in vivo studies have shown that overexpression of HSPB1 mutants in neurons is sufficient to cause pathological and electrophysiological changes typically observed in patients with motor neuropathy [86].

3.2. DnaJ(Hsp40)

Mutations in members of the Hsp40(DnaJ) family have been associated with neurologic and muscular disorders. Two examples are variants of the DNJAB2 and DNAJC5 genes, encoding members of the subfamilies B and C, respectively. DNAJB2 is mainly expressed in neurons and plays anti-protein aggregation and neuroprotective roles, as shown in models of neurodegenerative disorders [87,88,89]. The mutation c.352+1G-A in DNAJB2 was found associated with autosomal recessive distal spinal muscular atrophy-5 (DSMA5), a type of dHMN with young adult onset and characterized by progressive distal muscle weakness and atrophy with gait impairment and loss of reflexes [47] (Table 1). This genetic variant reduced or abolished chaperone expression, with a consequent accumulation and aggregation of misfolded proteins, which would cause lower motor neurons degeneration [47]. The same DNAJB2 variant was found in patients diagnosed with CMT2, and careful examination of phenotype and clinical evolution showed that pure motor impairment occurred early in the disease course, with no initial sensory symptoms, but sensory disturbances in the lower limbs appeared as the disease progressed [90]. Two other DNAJB2 gene variants were found in two families (Table 1). One (c.229+1G-A) showed a dHMN phenotype, with signs of distal muscular atrophy and paresis; the other, causing the missense mutation p.(Tyr5Cys), showed a CMT2 phenotype, with a noticeable sensory involvement, extending the group of DNAJB2-related diseases to include sensory neuropathy [46]. However, another evaluation of the phenotypes associated with these two latter genetic variations has classified them as DSMA5 (for details see the corresponding page on ClinVar database).
DNAJC5, also known as cysteine string protein (CSP), is abundant in neuronal cells and is a key element of the synaptic molecular machinery [48]. Two different genetic variations, an amino acid deletion and an amino acid substitution, were found associated with autosomal dominant adult-onset neuronal ceroid lipofuscinosis-4B (CLN4B), also known as Kufs’ disease [48,49,50,51]. These mutations affect two evolutionarily conserved leucine residues located in a conserved region of the cysteine-string domain that is involved in palmitoylation and membrane targeting of the protein [91]. CLN4B belongs to the group of neuronal ceroid lipofuscinosis (NCLs), a genetically heterogeneous group of at least nine neurodegenerative disorders, clinically characterized by progressive cognitive and motor impairment, visual impairment, epileptic seizures, and premature death. Despite the different ages of onset, all forms presented a common feature, i.e., the lysosomal accumulation of auto-fluorescent lipo-pigment in neuronal cells and peripheral tissues, which caused progressive and selective neurodegeneration and gliosis with secondary white matter lesions [92,93]. To clarify the molecular mechanisms responsible for the negative effects of the identified CSP mutations in neuronal cells, functional in vitro studies were performed. It was shown that both mutants had an abnormal intracellular localization and were less efficiently palmitoylated compared to the wild type protein. Moreover, they formed aggregates, causing CSP depletion that, in turn, could be responsible for the inhibition of synapse formation and synaptic transmission [49,94].

3.3. Hsp70(DnaK)

Pathogenic mutations in one of the members of the Hsp70(DnaK) family have been reported [26]. This chaperone family is composed of at least 17 members [95] but thus far only one of them, the mitochondrial mtHsp70, also named mortalin or HSPA9, has been found mutated and causing disease. Two mutations of this chaperone were associated with a syndrome very similar to CODAS (Cerebral, Ocular, Dental, Auricular, Skeletal Syndrome), which is caused by a mutation in the LONP1 gene, however it differs in that it also shows severe microtia, nasal hypoplasia, and other malformations. These characteristics are indicated by the name given to this new condition: EVEN-PLUS (from Epiphyseal, Vertebral, Ear, Nose, Plus associated findings) syndrome.

3.4. The Chaperonins

The best-known examples of genetic neurochaperonopathies are those involving members belonging to the family of chaperonins, including the group I chaperonin Hsp60 (HSPD1, of which a few pathogenic variants have been identified in humans) and the group II chaperonin CCT.
Chaperonins of Group I. The chaperonin HSPD1 is typically expressed within the mitochondrion where it assists the folding of proteins destined to the matrix together with the co-chaperonin Hsp10 (HSPE1) [45,96,97]. HSPD1 plays a key role not only in the maintenance of protein homeostasis in mitochondria, but also in sustaining cellular viability, since complete loss of the protein impaired mammalian development and postnatal survival, as demonstrated by an in vivo model of mice homozygous for an inactivation of HSPD1 gene [98]. In humans, missense mutations of the HSPD1 gene have been found associated with severe and chronic nervous system diseases. These include: 1) a dominantly inherited form of Spastic Paraplegia (SPG13) affecting motor neurons with the longest axons in the spinal cord and caused by either one of two missense mutations occurring at different positions, i.e., p.(Val98Ile) and p.(Gln461Glu), [54,55,99] (Table 1); and 2) a recessively inherited hypo-myelinating leukodystrophy (HLD4), also known as MitCHAP60 disease, which is a fatal early-onset neurodegenerative disorder characterized by pronounced cerebral hypomyelination and is caused by a single specific missense mutation, i.e., p.(Asp29Gly) [52,53] (Table 1). In vitro and in vivo studies have shown that these mutations destabilized the chaperonin oligomeric complex and reduced its ATPase and folding activities compared to the wild type form [100,101]. In addition to these two well-known neurological conditions, many others of various degrees of severity may be caused by other HSPD1 mutations that are known to occur in human genomes [97,102].
A point missense mutation, p.(Leu73Phe), in the gene encoding the co-chaperonin Hsp10 or HSPE1 (hsp10 or HSPE1 gene), has been found in a patient with a history of infantile spasms, hypotonia, developmental delay, a slightly enlarged liver, macrocephaly, and mild non-specific dysmorphic features [45] (Table 1).
The association between neurological disorders and HSPD1/E1 genes mutations is related to the great sensitivity of neuronal cells to mitochondrial dysfunction, since they have a metabolism mainly based on oxidative phosphorylation [103]. In vitro studies have shown that the expression of HSPD1 variants alters mitochondrial morphology, dynamics, and functionality [104,105].
Chaperonins of Group II. This group includes the CCT family, which in humans is composed of nine subunits: 1 through 8, with two versions of subunit 6; plus, three chaperones named BBS6, BBS10, and BBS12, which are involved in cilia biogenesis; and two other evolutionarily related molecules, CCT8L1 and CCT8L2, of which little is known [106] (Figure 1).
Here, we will consider chaperonopathies of the subunit CCT5 and those of the BBS cluster. BBS stands for Bardet–Biedl syndrome, which is the pathological condition caused by abnormalities in any one of these three chaperones. BBS6 is also called MKKS, because the disease associated with mutations in this gene is also called McKusick–Kaufmann syndrome.
The CCT subunits form functional hexadecamers with two octameric rings joined by their openings, which results in a barrel-like structure with a central cavity inside which polypeptide folding occurs [15]. This complex is also named TRiC, for TCP-1 ring complex, where TCP-1 stands for Tailless Complex Polypeptide 1. The integrity and functionality of this complicated multimolecular machine greatly depends on the integrity and functionality of all its individual components, the CCT subunits. For instance, a mutation in one of them may disrupt ring formation and hexadecamer assemblage or make the assemblage unstable and prone to dissociation, especially under the action of stressors.
A mutation occurring in subunit 5 of CCT was found to be associated with mutilating, distal sensory neuropathy [56,57] (Table 1). The functional CCT complex is composed of two hetero-octameric rings associated in a hexadecamer. The reported mutation, p.(His147Arg), occurred in subunit 5 (CCT5), and the disease was characterized by severe atrophy of the posterior tract of the spinal cord [56,57]. In vitro studies with human and archaeal purified molecules showed that the mutation, which occurs in the equatorial domain of the chaperonin subunit, impairs some of its properties and functions, resulting in poor oligomeric assembly [107,108,109].
More recently, a new disease was described associated with a different missense mutation, p.(Leu224Val), in CCT5 [58]. While the previously described mutation His147Arg caused a distal sensory neuropathy more pronounced in adulthood, the newly found mutation is associated with a motor neuropathy of early onset. Noteworthy is the different location of the mutations: His147Arg is in the equatorial domain, whereas Leu224Val is in the intermediate domain and appears to have an impact on the conformation of the apical domain, which is involved in substrate recognition and binding, namely functions quite distinct from those of the equatorial domain.
The disorders caused by mutations in the BBS genes affect ciliogenesis. Cilia are built with the participation of the BBSsome, an octameric complex composed of seven BBS proteins—BBS1, BBS2, BBS4, BBS5, BBS7, BBS8, and BBS9—and the protein BBIP10 [110]. BBS6, BBS10, and BBS12 serve as chaperones for BBSome assembly. Defects in the process of ciliogenesis and, consequently, in cilia structure and/or function, lead to a group of disorders called ciliopathies [24,111]. The Bardet–Biedl syndrome is a genetic disorder with autosomal recessive inheritance, highly prevalent in inbred/consanguineous populations, and with numerous primary and secondary clinical manifestations, including obesity, retinal degeneration, olfactory deficits, nephropathy, polydactyly, development delay, and cognitive or other neurological impairments [24,112]. About 70–80% of all cases of Bardet–Biedl syndrome result from mutations in BBS1 to BBS18 genes, with frequent mutations in BBS1 and BBS10 genes, especially in populations of European and Caucasian descent [113].
The BBS proteins excluded from the BBSome, i.e., BBS6, BBS10, and BBS12, have been classified as chaperonins, since they have high sequence similarity with members of the CCT family [63,65,106,114] (Figure 1). Even if these proteins are not components of the BBSome, mutations in their encoding genes also lead to the same phenotype resulting from the loss or abnormalities of BBSome subunits [24,65,115]. Moreover, BBS patients bearing pathogenic variants in the BBS6, BBS10, and BBS12 genes develop a more severe phenotype, with earlier disease onset and greater prevalence of all BBS primary diagnostic features typical of ciliopathies than patients with variants only in the non-chaperonin BBS genes [24,116]. The reason for this increased pathogenic severity associated with variants of the BBS chaperonin genes could be the malfunctioning of these three proteins, which are critical for BBSome assembly [24,117].
The Bardet–Biedl syndrome caused by mutations in the genes encoding for the chaperonin-like BBS proteins can be included in the group of genetic chaperonopathies [24], and in particular within the neurochaperonopathies if we consider the associated neurological abnormalities. In a mouse model of BBS, the morphological evaluation of brain neuroanatomy revealed ventriculomegaly of the lateral and third ventricles, thinning of the cerebral cortex, and reduced volume of the corpus striatum and hippocampus [118]. These abnormalities could be related to defects in the cilia of the ependymal cells, affecting cilia assembly, structure, and/or function [118,119].
Numerous genetic mutations in the genes encoding BBS6, BBS10, and BBS12 have been found [24,59,60,61,62,63,64,65,120]. Illustrative examples are listed in Table 1. Mutations in the BBS6 gene have been associated with typical BBS and with another similar disorder, the McKusick–Kaufman syndrome, and it was suggested that both syndromes are different allelic forms of the same clinical entity [59,121,122]. Moreover, in different families affected by BBS, many of the genetic variants in the BBS6 gene were found in the same individual [61] or associated with mutations in genes encoding BBS2 and other BBS proteins, suggesting a triallelic inheritance model for penetrance of the BBS phenotype [60,120]. BBS10 is, together with BBS1, the major contributor to BBS [113]. The most common mutation in the BBS10 is the 271dupT, occurring in 46% of mutant alleles [63].
Many of the BBS10 genetic variants were found in the same individual [62,64] or associated with mutations in the BBS1 gene [63]. Variants of BBS12 account for 8%–11% of the total cases of BBS in most of the cohorts reported [115].

4. Acquired Neurochaperonopathies

4.1. Hsp70(DnaK)

The Hsp70 family is composed of at least 17 members [95], and they are targets of a variety of PTMs such as phosphorylation, acetylation, ubiquitination, AMPylation, and ADP-ribosylation [123]. The only one human Hsp70 AMPylator enzyme containing Fic (Filamentation induced by cAMP) domain is the FICD enzyme, also known as Huntingtin yeast partner E (HYPE), able to catalyze the transfer of AMP onto a serine, threonine, or tyrosine residue of the substrate protein [124]. The consequences of Hsp70 AMPylation are poorly understood and, although there is no information on its impact on neurodegeneration to date, some interest in this regard has developed because of observations made using Saccharomyces cerevisiae, a eukaryote free of endogenous AMPylation machinery, and Caenorhabditis elegans models manipulated for the FIC-1 enzyme orthologous of human FICD (or HYPE) [125]. Induction of C. elegans FIC-1 (E274G) and Homo sapiens HYPE (E234G) in S. cerevisiae, in absence of stress, inhibited Hsp70 activity leading to decreased cell growth, toxic protein aggregation, and misfolding. These results suggest that AMPylation somehow modulates the Hsp70 protein folding machinery. Furthermore, other experiments confirmed AMPylation of a human Hsp70 family member, HSPA5, in a FICD-dependent manner, indicating that a high number of neuronal proteins involved in neuronal differentiation are AMPylated by FICD enzyme [126]. In addition, the fly Drosophila’s dFic AMPylator was shown to be essential for maintaining the levels of AMPylated Hsp70 BiP (the Hsp70 family member that resides in the endoplasmic reticulum (ER)), which are crucial for the photoreceptor’s resistance against stress; flies deprived of dFic are blind and unable of postsynaptic activation [123].
On the other hand, C. elegans models for Alzheimer’s disease (AD) and Parkinson ’s disease (PD) showed the formation of large protein aggregates when AMPylation of three Hsp70 family members was FIC-1 induced [127] (Table 2). Strikingly, the larger aggregates were more strongly cytoprotective than the smaller ones, extending the nematode’s life.

4.2. Hsp90 (HSPC)

The human Hsp90 (HSPC) family is composed of various members: two are cytosolic, one resides in the ER, and the other in the mitochondrion (references in [21]). Hsp90 chaperones have multiple sites for phosphorylation, acetylation, SUMOylation, methylation, O-GlcNAcylation, ubiquitination, and other PTMs [128]. Nitration of tyrosine has been associated with apoptosis of motor neurons in amyotrophic lateral sclerosis (ALS) patients and superoxide dismutase (SOD)-transgenic animal models [129] (Table 2). Nitration of five of the 24 tyrosine residues of HSPC2 and HSPC3 impaired ATPase chaperone activity. Nitrated tyrosine at position 56 of HSPC3 was identified in spinal cord sections of an ALS patient and of a mutated mouse model. The nitrated HSPC induced cell apoptosis by activation of P2 × 7 receptor-mediated calcium influx which induced Fas pathway in PC12 cells [129]. Several chronic and acute conditions causing oxidative stress induced the nitration of Hsp90 family members. However, the impact of Hsp90 nitration depends on the position of the nitrated tyrosine. In PC12 cells and motor neurons from rat embryo, it was observed that nitration of tyrosine at position 56 causes a gain of function, which is cytotoxic, whereas nitration of tyrosine at position 33 impairs mitochondrial activity [131].
Hsp90 collaborates with the Cell Division Cycle 37 (Cdc37) co-chaperone, forming a HSPC/Cdc37 complex, which binds almost 60% of the kinome preventing aggregation of activated regulators [132]. The co-chaperone Cdc37 is also a target of PTMs [133]. Phosphorylation at serine 13 of Cdc37 by CK2 (casein kinase II) affects the Hsp90/Cdc37 complex in performing its kinases homeostasis activity. In addition, the casein kinases 1 and 2 which involved AD, PD, and ALS pathogenesis have been reported to impair the interaction between Hsp90 and Cdc37 [134,135].

4.3. VCP/p97

The valosin-containing protein (p97 or VCP) is a homohexameric AAA+ ATPase protein of about 97 kDA. It is an abundant and ubiquitous molecule involved in cell cycle regulation [136], autophagy [137], DNA repair [138], and ER-associated and ubiquitin-mediated protein degradation [139]. VCP has been found within protein aggregates in PD, ALS, AD, and in polyglutamine aggregates [130]. Polyglutamine disease is characterized by deacetylation of histones H3 and H4 and a correlation of this deacetylation with VCP activity has been suggested [130]. VCP was found in the PC12 cell nucleus in protein aggregates with polyglutamine expansions, showing PTM, such as phosphorylation at Ser-612 and Thr-613, and acetylation at Lys-614 [130]. Nuclear, post-translationally modified VCP is associated with neurite retraction and shrinkage, features which have been found also in PC12 cells expressing protein aggregates with polyglutamine expansions [130]. Therefore, it has been suggested that VCP could be a mediator of histones H3 and H4 deacetylation, even if the underlined mechanism remains to be elucidated. VCP appears to be essential for the clearance of “occasionally” aggregated proteins, promoting the deacetylation of H3 and H4 and, in turn, reducing the transcription of de novo proteins; however, during chronic and prolonged protein aggregation, nuclear translocation and post-translational modification of VCP could induce neurodegeneration and cause an atrophic phenotype as observed in the Drosophila experimental model [130].

5. Discussion

In the last few years, an increasing number of scientific reports have highlighted the involvement of members of the CS in the pathogenesis of diverse neurological disorders. It was found that genetic and/or acquired malfunction (or lack of function) of members of the CS are implicated in the pathogenic mechanisms underlying NCPs (Figure 2).
The heterogeneity of these disorders and the difficulties implied in the study of the nervous system in humans are the major challenges for physicians and other investigators willing to elucidate the relevant molecular mechanisms and to develop efficacious treatment strategies. The data reported thus far, some of which has been briefly discussed in the previous sections, indicate that pathogenic variants of members of the sHsp, DnaJ(Hsp40), and CCT families, and of the Hsp60 gene, tend to cause predominantly motor disorders, although sensory and mental functions may also be affected [24,28,32,33,36,46,49,52,54]. For example, a missense mutation of the subunit number 5 of CCT, CCT5, was reported to be accompanied by a phenotype characterized mostly by distal sensory mutilating neuropathy [56,57]. In contrast, another recently discovered mutant located in a different structural domain of CCT5 was associated with a motor distal neuropathy without mutilation [58]. Thus, the impact of the mutation on the properties and functions of the chaperone molecule, and the accompanying tissue abnormalities and lesions observed in patients may greatly differ, depending on the location of the mutant amino acid, but other elements, e.g., environmental and nutritional factors that may also contribute to the generation of diverse phenotypes cannot be excluded.
Other genetic NCPs discussed are caused by variants of a subfamily of CCT genes composed of the BBS6 (MKKS), BBS10, and BBS12 genes [24,59,60,61,63,64,65]. Cognitive deficit was recorded only in patients with pathogenic variants of the CCT5 (the newly reported mutation p.(Leu224Val)), DNAJC, and BBS genes (Figure 2).
Only one of the members of the Hsp70(DnaK) family, the mitochondrial Hsp70 (mtHsp70 also named mortalin or HSPA9) has been reported with pathogenic mutations causing a syndrome that can be classified as NCP. Other members of this family, such as the HSPA1A and HSPA1B, whose genes are organized in tandem within the so called HSPA1-cluster [140] are stress-inducible and have been implicated in the pathogenesis of several diseases, such as cancer and neurodegenerative conditions, but their genes were normal [141,142]. This pathogenicity of the Hsp70 molecules fits the description of acquired chaperonopathies by mistake or collaborationism, in which apparently normal chaperones help the disease rather than protect from it. The lack of genetic NCPs associated with mutations in HSPA1A and HSPA1B genes could be related to their vital roles suggested also by their evolutionary conservation [95]. This is also suggested by the fact that the majority of polymorphisms identified in the human HSPA1-cluster are in the UTR regions, and the most common SNPs occurring in the coding regions are synonymous and thus, do not affect the proteins’ amino acid sequences [143,144]. A few studies experimentally investigated the impact of human HSPA1A-1B mutations on the function/expression of the encoded proteins [145,146,147,148]. The functional analysis of six naturally occurring variants of the HSPA1A gene revealed that these variants result in significant functional alterations, affecting ATP-hydrolysis rate and substrate-binding ability, likely by altering the allosteric communication between its two major domains. However, they are rare mutations, with very low penetrance in the human population. All these results support the notion that the HSPA1A gene is under strong selection and that its functions must be conserved [149], which would explain why no patients with pathogenic mutations of this gene have been found.
The Hsp90 (HSPC) family, including five members and various isoforms expressed in specific cellular organelles, is highly conserved from early eukaryotes to humans and plays important roles in protein homeostasis (references in [21]). Hsp90 chaperones have a wide array of client proteins and, together with their co-chaperones and co-factors, participate in cell survival, cell-cycle control mechanisms, hormone signaling and other signaling pathways, and apoptosis [150]. In addition, Hsp90 has been implicated in the pathogenesis of Alzheimer’s disease and Parkinson’s disease [151,152]. However, it is noteworthy that no Hsp90 mutations have been described yet as an etiological-pathogenic factor in NCPs. The same message is given by analysis of Hsp90 genetic polymorphisms. Polymorphic variants in introns and exons of the human Hsp90beta (HSPC3) gene have been detected with no apparent effects on phenotype [153].
In recent years, along with genetic NCPs, attention has also been directed to acquired NCPs associated with molecular chaperones with PTMs. For instance, in vivo studies demonstrated that changes in the AMPylation balance of molecular chaperones could influence cell fate in physiological [125] and pathological conditions, such as PD and AD [127]. Human FICD AMPylates various amino acids in the Hsp70, DNAJB1, and Hsp90 chaperones. AMPylation levels regulate the correct neurogenesis and neuronal conduction [123,126]. So far there are no data suggesting that chaperone AMPylation may be disruptive and contribute to the pathogenic mechanism of neurological disorders.
Nitration of Hsp90 (HSPC) family members in relation to ALS disease is an example of the differential outcomes one can expect depending on the site and level of the PTM. Furthermore, considering the importance of mitochondrial activity in neuronal cells, one may assume that nitrated Hsp90 may be associated with acquired NCPs in addition to those for ALS. Acquired NCPs could be the result of PTMs in molecular chaperones and of PTMs in co-chaperones, chaperone co-factors or chaperone interactors, as illustrated by the consequences of phosphorylation of Hsp90/Cdc37 complex [133,135].
Two coexisting PTMs, phosphorylation and acetylation, in the VCP in the nucleus may promote histone deacetylation [138]. VCP could also undergo methylation and S-glutathionylation, two modifications that occur during oxidative stress with impairment of ATPase activity [139]. Since oxidative stress is a prevalent feature of neurodegenerative diseases, it is possible that modified VCP also plays a pathogenic role in NCPs.
Pathogenic chaperones with aberrant PTMs associated with NCPs, other than those described above, have not been reported. Most of the naturally occurring or induced PTMs in Hsp60 (HSPD1) with a pathogenic role have been implicated in carcinogenesis [154,155,156,157], apoptosis signaling regulation [158,159], and immune system regulation [160], but none has been associated with NCPs [25,161]. However, an involvement of Hsp60 (HSPD1) in AD and PD is under scrutiny. It is possible that Hsp60 with aberrant PTMs and/or with abnormal levels and location, i.e., intra- or extracellular, and when intracellular intra- or extramitochondrial, may be pathogenic in neurological disorders [25].
The number of multiple PTMs that can occur on a molecular chaperone is strikingly high, e.g., 60% and 56% of the T, Y, and S residues of Hsp70 (HSPA8) and Hsp90, respectively, could be phosphorylated [123] (Figure 3).
In addition, the same residue could be subjected to different PTMs, and several PTMs can also occur on the relevant co-chaperones and collaborators. All these modifications create the “Chaperone Code” that regulates chaperone functions, physiologically and pathologically (e.g., in cancer, viral infections, chronic inflammation, and neurodegenerative disorders). This flexibility and diversity of potential functions in health and disease of a single chaperone type, together with the variety of chaperone types, opens exciting novel avenues for investigating the mechanisms involved in acquired NCPs, for example those characteristics of ageing and AD and PD [135].

6. Conclusions and Future Perspective

Why is it worthwhile to determine if any given neurological patient bears a chaperonopathy? Firstly, the CS is present in all cells, tissues, and organs, therefore, a deficiency in any of its components may have a pathogenic impact in various anatomic areas and physiological systems, including the nervous system, and the probability of causing signs and symptoms is high. Secondly, the CS participates in a variety of physiological mechanisms, both as guardian of protein homeostasis and protector against stressors, and as the effector of many other cellular processes unrelated to protein quality control, i.e., non-canonical functions, including critical interactions with the immune system. This also increases the probability of deficiencies in the CS having clinical manifestations. Thirdly, detection of a chaperonopathy opens the door to investigate the possibility of applying chaperonotherapy in any of its modes, which is a significant prospect considering that most chaperonopathies are serious diseases still awaiting efficacious treatment [20,164,165,166,167]. It would be beneficial for many patients if physicians would bear in mind the concept of hidden chaperonopathies, particularly in those cases in which the clinical signs and symptoms do not quite fit within the expected clinical picture for any given neurological disorder [168]. To unveil this hidden pathogenic factor, physicians should add a search for a chaperonopathy to the differential diagnosis algorithm.

Author Contributions

Project planning, F.C. and F.S.; literature searches, A.M.V., R.S., F.S., E.C.d.M. and A.J.L.M.; molecular modeling, F.S. and A.M.V.; writing, F.S., A.M.V., R.S., E.C.d.M. and A.J.L.M. All authors have read and agreed to the published version of the manuscript.

Funding

A.J.L.M and E.C.d.M. were partially supported by IMET and IEMEST. This is IMET contribution number IMET 21-002.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Briston, T.; Hicks, A.R. Mitochondrial dysfunction and neurodegenerative proteinopathies: Mechanisms and prospects for therapeutic intervention. Biochem. Soc. Trans. 2018, 46, 829–842. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Dugger, B.N.; Dickson, D.W. Pathology of neurodegenerative diseases. Cold Spring Harb. Perspect. Biol. 2017, 9, a028035. [Google Scholar] [CrossRef] [PubMed]
  3. Gao, J.; Wang, L.; Yan, T.; Perry, G.; Wang, X. TDP-43 proteinopathy and mitochondrial abnormalities in neurodegeneration. Mol. Cell. Neurosci. 2019, 100, 103396. [Google Scholar] [CrossRef] [PubMed]
  4. Robinson, J.L.; Lee, E.B.; Xie, S.X.; Rennert, L.; Suh, E.; Bredenberg, C.; Caswell, C.; Van Deerlin, V.M.; Yan, N.; Yousef, A.; et al. Neurodegenerative disease concomitant proteinopathies are prevalent, age-related and APOE4-associated. Brain 2018, 141, 2181–2193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Carlisle, C.; Prill, K.; Pilgrim, D. Chaperones and the proteasome system: Regulating the construction and demolition of striated muscle. Int. J. Mol. Sci. 2018, 19, 32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Dikic, I. Proteasomal and autophagic degradation systems. Annu. Rev. Biochem. 2017, 86, 193–224. [Google Scholar] [CrossRef]
  7. Nam, T.; Han, J.H.; Devkota, S.; Lee, H.W. Emerging paradigm of crosstalk between autophagy and the ubiquitin-proteasome system. Mol. Cells 2017, 40, 897–905. [Google Scholar]
  8. Catarino, S.; Pereira, P.; Girao, H. Molecular control of chaperone-mediated autophagy. Essays Biochem. 2017, 61, 663–674. [Google Scholar]
  9. Tekirdag, K.; Cuervo, A.M. Chaperone-mediated autophagy and endosomal microautophagy: Joint by a chaperone. J. Biol. Chem. 2018, 293, 5414–5424. [Google Scholar] [CrossRef] [Green Version]
  10. Kocaturk, N.M.; Gozuacik, D. Crosstalk between mammalian autophagy and the ubiquitin-proteasome system. Front. Cell Dev. Biol. 2018, 6, 128. [Google Scholar] [CrossRef]
  11. Horwich, A.L.; Fenton, W.A.; Chapman, E.; Farr, G.W. Two families of chaperonin: Physiology and mechanism. Annu. Rev. Cell Dev. Biol. 2007, 23, 115–145. [Google Scholar] [CrossRef] [PubMed]
  12. Kim, Y.E.; Hipp, M.S.; Bracher, A.; Hayer-Hartl, M.; Ulrich Hartl, F. Molecular Chaperone Functions in Protein Folding and Proteostasis. Annu. Rev. Biochem. 2013, 82, 323–355. [Google Scholar] [CrossRef] [PubMed]
  13. Finka, A.; Sharma, S.K.; Goloubinoff, P. Multi-layered molecular mechanisms of polypeptide holding, unfolding and disaggregation by HSP70/HSP110 chaperones. Front. Mol. Biosci. 2015, 2, 29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Mogk, A.; Bukau, B.; Kampinga, H.H. Cellular Handling of Protein Aggregates by Disaggregation Machines. Mol. Cell 2018, 69, 214–226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Willison, K.R. The structure and evolution of eukaryotic chaperonin-containing TCP-1 and its mechanism that folds actin into a protein spring. Biochem. J. 2018, 475, 3009–3034. [Google Scholar] [CrossRef]
  16. Adams, B.M.; Oster, M.E.; Hebert, D.N. Protein Quality Control in the Endoplasmic Reticulum. Protein J. 2019, 38, 317–329. [Google Scholar] [CrossRef]
  17. Dahiya, V.; Buchner, J. Functional principles and regulation of molecular chaperones. In Advances in Protein Chemistry and Structural Biology; Academic Press Inc.: Cambridge, MA, USA, 2019; Volume 114, pp. 1–60. ISBN 9780128155578. [Google Scholar]
  18. Macario, A.J.L.; Conway de Macario, E. Sick chaperones, cellular stress, and disease. N. Engl. J. Med. 2005, 353, 1489–1501. [Google Scholar] [CrossRef]
  19. Macario, A.J.L.; Conway de Macario, E. The chaperoning system: Physiology and pathology. In Experimental Medicine Reviews; Plumelia: Bagheria, Italy, 2008; Volumes 2–3, pp. 9–21. [Google Scholar]
  20. Macario, A.J.L.; Conway de Macario, E. Chaperone proteins and chaperonopathies. In Handbook of Stress; Elsevier; Academic Press: Cambridge, MS, USA, 2019. [Google Scholar]
  21. Macario, A.J.L.; Conway de Macario, E.; Cappello, F. The Chaperonopathies. Diseases with Defective Molecular Chaperones; Springer: Berlin, Germany, 2013; ISBN 978-94-007-4666-4. [Google Scholar]
  22. Lanfranco, M.; Vassallo, N.; Cauchi, R.J. Spinal muscular atrophy: From defective chaperoning of snRNP assembly to neuromuscular dysfunction. Front. Mol. Biosci. 2017, 4, 41. [Google Scholar] [CrossRef] [Green Version]
  23. Lupo, V.; Aguado, C.; Knecht, E.; Espinós, C. Chaperonopathies: Spotlight on hereditary motor neuropathies. Front. Mol. Biosci. 2016, 3, 81. [Google Scholar] [CrossRef] [Green Version]
  24. Álvarez-Satta, M.; Castro-Sánchez, S.; Valverde, D. Bardet-Biedl syndrome as a chaperonopathy: Dissecting the major role of chaperonin-like BBS proteins (BBS6-BBS10-BBS12). Front. Mol. Biosci. 2017, 4, 55. [Google Scholar] [CrossRef] [Green Version]
  25. Caruso Bavisotto, C.; Alberti, G.; Vitale, A.M.; Paladino, L.; Campanella, C.; Rappa, F.; Gorska, M.; Conway de Macario, E.; Cappello, F.; Macario, A.J.L.; et al. Hsp60 Post-translational Modifications: Functional and Pathological Consequences. Front. Mol. Biosci. 2020, 7, 95. [Google Scholar] [CrossRef] [PubMed]
  26. Royer-Bertrand, B.; Castillo-Taucher, S.; Moreno-Salinas, R.; Cho, T.J.; Chae, J.H.; Choi, M.; Kim, O.H.; Dikoglu, E.; Campos-Xavier, B.; Girardi, E.; et al. Mutations in the heat-shock protein A9 (HSPA9) gene cause the EVEN-PLUS syndrome of congenital malformations and skeletal dysplasia. Sci. Rep. 2015, 5, 17154. [Google Scholar] [CrossRef] [PubMed]
  27. Ghaoui, R.; Palmio, J.; Brewer, J.; Lek, M.; Needham, M.; Evilä, A.; Hackman, P.; Jonson, P.H.; Penttilä, S.; Vihola, A.; et al. Mutations in HSPB8 causing a new phenotype of distal myopathy and motor neuropathy. Neurology 2016, 86, 391–398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Irobi, J.; Van Impe, K.; Seeman, P.; Jordanova, A.; Dierick, I.; Verpoorten, N.; Michalik, A.; De Vriendt, E.; Jacobs, A.; Van Gerwen, V.; et al. Hot-spot residue in small heat-shock protein 22 causes distal motor neuropathy. Nat. Genet. 2004, 36, 597–601. [Google Scholar] [CrossRef] [Green Version]
  29. Timmerman, V.; Raeymaekers, P.; Nelis, E.; De Jonghe, P.; Muylle, L.; Ceuterick, C.; Martin, J.J.; Van Broeckhoven, C. Linkage analysis of distal hereditary motor neuropathy type II (distal HMN II) in a single pedigree. J. Neurol. Sci. 1992, 109, 41–48. [Google Scholar] [CrossRef]
  30. Tang, B.S.; Luo, W.; Xia, K.; Xiao, J.F.; Jiang, H.; Shen, L.; Tang, J.G.; Zhao, G.H.; Cai, F.; Pan, Q.; et al. A new locus for autosomal dominant Charcot-Marie-Tooth disease type 2 (CMT2L) maps to chromosome 12q. Hum. Genet. 2004, 114, 527–533. [Google Scholar]
  31. Tang, B.S.; Zhao, G.; Luo, W.; Xia, K.; Cai, F.; Pan, Q.; Zhang, R.; Zhang, F.; Liu, X.; Chen, B.; et al. Small heat-shock protein 22 mutated in autosomal dominant Charcot-Marie-Tooth disease type 2L. Hum. Genet. 2005, 116, 222–224. [Google Scholar] [CrossRef]
  32. Capponi, S.; Geroldi, A.; Fossa, P.; Grandis, M.; Ciotti, P.; Gulli, R.; Schenone, A.; Mandich, P.; Bellone, E. HSPB1 and HSPB8 in inherited neuropathies: Study of an Italian cohort of dHMN and CMT2 patients. J. Peripher. Nerv. Syst. 2011, 16, 287–294. [Google Scholar] [CrossRef]
  33. Houlden, H.; Laura, M.; Wavrant-De Vrièze, F.; Blake, J.; Wood, N.; Reilly, M.M. Mutations in the HSP27 (HSPB1) gene cause dominant, recessive, and sporadic distal HMN/CMT type. Neurology 2008, 71, 1660–1668. [Google Scholar] [CrossRef]
  34. James, P.A.; Rankin, J.; Talbot, K. Asymmetrical late onset motor neuropathy associated with a novel mutation in the small heat shock protein HSPB1 (HSP27). J. Neurol. Neurosurg. Psychiatry 2008, 79, 461–463. [Google Scholar] [CrossRef]
  35. Ho, C.C.; Tai, S.M.; Lee, E.C.N.; Mak, T.S.H.; Liu, T.K.T.; Tang, V.W.L.; Poon, W.T. Rapid identification of pathogenic variants in two cases of Charcot-Marie-tooth disease by gene-panel sequencing. Int. J. Mol. Sci. 2017, 18, 770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Evgrafov, O.V.; Mersiyanova, I.; Irobi, J.; Van Den Bosch, L.V.; Dierick, I.; Leung, C.L.; Schagina, O.; Verpoorten, N.; Van Impe, K.; Fedotov, V.; et al. Mutant small heat-shock protein 27 causes axonal Charcot-Marie-Tooth disease and distal hereditary motor neuropathy. Nat. Genet. 2004, 36, 602–606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Tang, B.; Liu, X.; Zhao, G.; Luo, W.; Xia, K.; Pan, Q.; Cai, F.; Hu, Z.; Zhang, C.; Chen, B.; et al. Mutation analysis of the small heat shock protein 27 gene in Chinese patients with Charcot-Marie-Tooth disease. Arch. Neurol. 2005, 62, 1201–1207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Benedetti, S.; Previtali, S.C.; Coviello, S.; Scarlato, M.; Cerri, F.; Di Pierri, E.; Piantoni, L.; Spiga, I.; Fazio, R.; Riva, N.; et al. Analyzing histopathological features of rare Charcot-Marie-Tooth neuropathies to unravel their pathogenesis. Arch. Neurol. 2010, 67, 1498–1505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Oberstadt, M.; Mitter, D.; Classen, J.; Baum, P. Late onset dHMN II caused by c.404C>G mutation in HSPB1 gene. J. Peripher. Nerv. Syst. 2016, 21, 111–113. [Google Scholar] [CrossRef]
  40. Ismailov, S.M.; Fedotov, V.P.; Dadali, E.L.; Polyakov, A.V.; Van Broeckhoven, C.; Ivanov, V.I.; De Jonghe, P.; Timmerman, V.; Evgrafov, O.V. A new locus for autosomal dominant Charcot-Marie-Tooth disease type 2 (CMT2F) maps to chromosome 7q11-q. Eur. J. Hum. Genet. 2001, 9, 646–650. [Google Scholar] [CrossRef] [Green Version]
  41. Ikeda, Y.; Abe, A.; Ishida, C.; Takahashi, K.; Hayasaka, K.; Yamada, M. A clinical phenotype of distal hereditary motor neuronopathy type II with a novel HSPB1 mutation. J. Neurol. Sci. 2009, 277, 9–12. [Google Scholar] [CrossRef]
  42. Lin, K.-P.; Soong, B.-W.; Yang, C.-C.; Huang, L.-W.; Chang, M.-H.; Lee, I.-H.; Antonellis, A.; Lee, Y.-C. The Mutational Spectrum in a Cohort of Charcot-Marie-Tooth Disease Type 2 among the Han Chinese in Taiwan. PLoS ONE 2011, 6, e29393. [Google Scholar] [CrossRef] [Green Version]
  43. Luigetti, M.; Fabrizi, G.M.; Madia, F.; Ferrarini, M.; Conte, A.; Del Grande, A.; Tasca, G.; Tonali, P.A.; Sabatelli, M. A novel HSPB1 mutation in an Italian patient with CMT2/dHMN phenotype. J. Neurol. Sci. 2010, 298, 114–117. [Google Scholar] [CrossRef]
  44. Kijima, K.; Numakura, C.; Goto, T.; Takahashi, T.; Otagiri, T.; Umetsu, K.; Hayasaka, K. Small heat shock protein 27 mutation in a Japanese patient with distal hereditary motor neuropathy. J. Hum. Genet. 2005, 50, 473–476. [Google Scholar] [CrossRef] [Green Version]
  45. Bie, A.S.; Fernandez-Guerra, P.; Birkler, R.I.D.; Nisemblat, S.; Pelnena, D.; Lu, X.; Deignan, J.L.; Lee, H.; Dorrani, N.; Corydon, T.J.; et al. Effects of a mutation in the HSPE1 gene encoding the mitochondrial co-chaperonin HSP10 and its potential association with a neurological and developmental disorder. Front. Mol. Biosci. 2016, 3, 65. [Google Scholar] [CrossRef] [PubMed]
  46. Gess, B.; Auer-Grumbach, M.; Schirmacher, A.; Strom, T.; Zitzelsberger, M.; Rudnik-Schöneborn, S.; Röhr, D.; Halfter, H.; Young, P.; Senderek, J. HSJ1-related hereditary neuropathies Novel mutations and extended clinical spectrum. Neurology 2014, 83, 1726–1732. [Google Scholar] [CrossRef] [PubMed]
  47. Blumen, S.C.; Astord, S.; Robin, V.; Vignaud, L.; Toumi, N.; Cieslik, A.; Achiron, A.; Carasso, R.L.; Gurevich, M.; Braverman, I.; et al. A rare recessive distal hereditary motor neuropathy with HSJ1 chaperone mutation. Ann. Neurol. 2012, 71, 509–519. [Google Scholar]
  48. Benitez, B.A.; Alvarado, D.; Cai, Y.; Mayo, K.; Chakraverty, S.; Norton, J.; Morris, J.C.; Sands, M.S.; Goate, A.; Cruchaga, C. Exome-sequencing confirms DNAJC5 mutations as cause of adult neuronal ceroid-lipofuscinosis. PLoS ONE 2011, 6, 26741. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Nosková, L.; Stránecký, V.; Hartmannová, H.; Přistoupilová, A.; Barešová, V.; Ivánek, R.; Hlková, H.; Jahnová, H.; Van Der Zee, J.; Staropoli, J.F.; et al. Mutations in DNAJC5, encoding cysteine-string protein alpha, cause autosomal-dominant adult-onset neuronal ceroid lipofuscinosis. Am. J. Hum. Genet. 2011, 89, 241–252. [Google Scholar] [CrossRef] [Green Version]
  50. Velinov, M.; Dolzhanskaya, N.; Gonzalez, M.; Powell, E.; Konidari, I.; Hulme, W.; Staropoli, J.F.; Xin, W.; Wen, G.Y.; Barone, R.; et al. Mutations in the Gene DNAJC5 Cause Autosomal Dominant Kufs Disease in a Proportion of Cases: Study of the Parry Family and 8 Other Families. PLoS ONE 2012, 7, e29729. [Google Scholar] [CrossRef]
  51. Cadieux-Dion, M.; Andermann, E.; Lachance-Touchette, P.; Ansorge, O.; Meloche, C.; Barnabé, A.; Kuzniecky, R.I.; Andermann, F.; Faught, E.; Leonberg, S.; et al. Recurrent mutations in DNAJC5 cause autosomal dominant Kufs disease. Clin. Genet. 2013, 83, 571–575. [Google Scholar] [CrossRef]
  52. Magen, D.; Georgopoulos, C.; Bross, P.; Ang, D.; Segev, Y.; Goldsher, D.; Nemirovski, A.; Shahar, E.; Ravid, S.; Luder, A.; et al. Mitochondrial Hsp60 Chaperonopathy Causes an Autosomal-Recessive Neurodegenerative Disorder Linked to Brain Hypomyelination and Leukodystrophy. Am. J. Hum. Genet. 2008, 83, 30–42. [Google Scholar] [CrossRef] [Green Version]
  53. Kusk, M.S.; Damgaard, B.; Risom, L.; Hansen, B.; Ostergaard, E. Hypomyelinating Leukodystrophy due to HSPD1 Mutations: A New Patient. Neuropediatrics 2016, 47, 332–335. [Google Scholar] [CrossRef]
  54. Hansen, J.J.; Dürr, A.; Cournu-Rebeix, I.; Georgopoulos, C.; Ang, D.; Nielsen, M.N.; Davoine, C.S.; Brice, A.; Fontaine, B.; Gregersen, N.; et al. Hereditary spastic paraplegia SPG13 is associated with a mutation in the gene encoding the mitochondrial chaperonin Hsp60. Am. J. Hum. Genet. 2002, 70, 1328–1332. [Google Scholar] [CrossRef] [Green Version]
  55. Hansen, J.; Svenstrup, K.; Ang, D.; Nielsen, M.N.; Christensen, J.H.; Gregersen, N.; Nielsen, J.E.; Georgopoulos, C.; Bross, P. A novel mutation in the HSPD1 gene in a patient with hereditary spastic paraplegia. J. Neurol. 2007, 254, 897–900. [Google Scholar] [CrossRef] [PubMed]
  56. Bouhouche, A.; Benomar, A.; Bouslam, N.; Chkili, T.; Yahyaoui, M. Mutation in the epsilon subunit of the cytosolic chaperonin-containing t-complex peptide-1 (Cct5) gene causes autosomal recessive mutilating sensory neuropathy with spastic paraplegia. J. Med. Genet. 2006, 43, 441–443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Bouhouche, A.; Benomar, A.; Bouslam, N.; Ouazzani, R.; Chkili, T.; Yahyaoui, M. Autosomal recessive mutilating sensory neuropathy with spastic paraplegia maps to chromosome 5p15.31-14. Eur. J. Hum. Genet. 2006, 14, 249–252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Antona, V.; Scalia, F.; Giorgio, E.; Radio, F.C.; Brusco, A.; Oliveri, M.; Corsello, G.; Lo Celso, F.; Vadalà, M.; Conway de Macario, E.; et al. A novel cct5 missense variant associated with early onset motor neuropathy. Int. J. Mol. Sci. 2020, 21, 7631. [Google Scholar] [CrossRef] [PubMed]
  59. Katsanis, N.; Beales, P.L.; Woods, M.O.; Lewis, R.A.; Green, J.S.; Parfrey, P.S.; Ansley, S.J.; Davidson, W.S.; Lupski, J.R. Mutations in MKKS cause obesity, retinal dystrophy and renal malformations associated with Bardet-Biedl syndrome. Nat. Genet. 2000, 26, 67–70. [Google Scholar] [CrossRef] [PubMed]
  60. Katsanis, N.; Lupski, J.R.; Beales, P.L. Exploring the molecular basis of Bardet-Biedl syndrome. Hum. Mol. Genet. 2001, 10, 2293–2299. [Google Scholar] [CrossRef] [PubMed]
  61. Slavotinek, A.M.; Stone, E.M.; Mykytyn, K.; Heckenlively, J.R.; Green, J.S.; Heon, E.; Musarella, M.A.; Parfrey, P.S.; Sheffield, V.C.; Biesecker, L.G. Mutations in MKKS cause Bardet-Biedl syndrome. Nat. Genet. 2000, 26, 15–16. [Google Scholar] [CrossRef]
  62. Laurier, V.; Stoetzel, C.; Muller, J.; Thibault, C.; Corbani, S.; Jalkh, N.; Salem, N.; Chouery, E.; Poch, O.; Licaire, S.; et al. Pitfalls of homozygosity mapping: An extended consanguineous Bardet-Biedl syndrome family with two mutant genes (BBS2, BBS10), three mutations, but no triallelism. Eur. J. Hum. Genet. 2006, 14, 1195–1203. [Google Scholar] [CrossRef]
  63. Stoetzel, C.; Laurier, V.; Davis, E.E.; Muller, J.; Rix, S.; Badano, J.L.; Leitch, C.C.; Salem, N.; Chouery, E.; Corbani, S.; et al. BBS10 encodes a vertebrate-specific chaperonin-like protein and is a major BBS locus. Nat. Genet. 2006, 38, 521–524. [Google Scholar] [CrossRef]
  64. Putoux, A.; Mougou-Zerelli, S.; Thomas, S.; Elkhartoufi, N.; Audollent, S.; Le Merrer, M.; Lachmeijer, A.; Sigaudy, S.; Buenerd, A.; Fernandez, C.; et al. BBS10 mutations are common in ’Meckel’-type cystic kidneys. J. Med. Genet. 2010, 47, 848–852. [Google Scholar] [CrossRef]
  65. Stoetzel, C.; Muller, J.; Laurier, V.; Davis, E.E.; Zaghloul, N.A.; Vicaire, S.; Jacquelin, C.; Plewniak, F.; Leitch, C.C.; Sarda, P.; et al. Identification of a novel BBS gene (BBS12) highlights the major role of a vertebrate-specific branch of chaperonin-related proteins in Bardet-Biedl syndrome. Am. J. Hum. Genet. 2007, 80, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Garrido, C.; Paul, C.; Seigneuric, R.; Kampinga, H.H. The small heat shock proteins family: The long forgotten chaperones. Int. J. Biochem. Cell Biol. 2012, 44, 1588–1592. [Google Scholar] [CrossRef] [PubMed]
  67. Bruey, J.M.; Ducasse, C.; Bonniaud, P.; Ravagnan, L.; Susin, S.A.; Diaz-Latoud, C.; Gurbuxani, S.; Arrigo, A.P.; Kroemer, G.; Solary, E.; et al. Hsp27 negatively regulates cell death by interacting with cytochrome c. Nat. Cell Biol. 2000, 2, 645–652. [Google Scholar] [CrossRef] [PubMed]
  68. Pandey, P.; Farber, R.; Nakazawa, A.; Kumar, S.; Bharti, A.; Nalin, C.; Weichselbaum, R.; Kufe, D.; Kharbanda, S. Hsp27 functions as a negative regulator of cytochrome c-dependent activation of procaspase. Oncogene 2000, 19, 1975–1981. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Chowdary, T.K.; Raman, B.; Ramakrishna, T.; Rao, C.M. Mammalian Hsp22 is a heat-inducible small heat-shock protein with chaperone-like activity. Biochem. J. 2004, 381, 379–387. [Google Scholar] [CrossRef] [Green Version]
  70. Vos, M.J.; Zijlstra, M.P.; Kanon, B.; van Waarde-Verhagen, M.A.W.H.; Brunt, E.R.P.; Oosterveld-Hut, H.M.J.; Carra, S.; Sibon, O.C.M.; Kampinga, H.H. HSPB7 is the most potent polyQ aggregation suppressor within the HSPB family of molecular chaperones. Hum. Mol. Genet. 2010, 19, 4677–4693. [Google Scholar] [CrossRef] [Green Version]
  71. Kawano, F.; Fujita, R.; Nakai, N.; Terada, M.; Ohira, T.; Ohira, Y. HSP25 can modulate myofibrillar desmin cytoskeleton following the phosphorylation at Ser15 in rat soleus muscle. J. Appl. Physiol. 2012, 112, 176–186. [Google Scholar] [CrossRef]
  72. Van Montfort, R.; Slingsby, C.; Vierling, E. Structure and function of the small heat shock protein/α-crystallin family of molecular chaperones. Adv. Protein Chem. 2001, 59, 105–156. [Google Scholar]
  73. Boncoraglio, A.; Minoia, M.; Carra, S. The family of mammalian small heat shock proteins (HSPBs): Implications in protein deposit diseases and motor neuropathies. Int. J. Biochem. Cell Biol. 2012, 44, 1657–1669. [Google Scholar] [CrossRef]
  74. Berger, P.; Young, P.; Suter, U. Molecular cell biology of Charcot-Marie-Tooth disease. Neurogenetics 2002, 4, 1–15. [Google Scholar] [CrossRef]
  75. Harding, A.E. Inherited Neuronal Atrophy and Degeneration Predominantly of Lower Motor Neurons. In Peripheral Neuropathy; Elsevier: Amsterdam, The Netherlands, 2005; Volume 2, pp. 1603–1621. ISBN 9780721694917. [Google Scholar]
  76. Rossor, A.M.; Kalmar, B.; Greensmith, L.; Reilly, M.M. The distal hereditary motor neuropathies. J. Neurol. Neurosurg. Psychiatry 2012, 83, 6–14. [Google Scholar] [CrossRef] [PubMed]
  77. Solla, P.; Vannelli, A.; Bolino, A.; Marrosu, G.; Coviello, S.; Murru, M.R.; Tranquilli, S.; Corongiu, D.; Benedetti, S.; Marrosu, M.G. Heat shock protein 27 R127W mutation: Evidence of a continuum between axonal Charcot-Marie-Tooth and distal hereditary motor neuropathy. J. Neurol. Neurosurg. Psychiatry 2010, 81, 958–962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Weeks, S.D.; Muranova, L.K.; Heirbaut, M.; Beelen, S.; Strelkov, S.V.; Gusev, N.B. Characterization of human small heat shock protein HSPB1 α-crystallin domain localized mutants associated with hereditary motor neuron diseases. Sci. Rep. 2018, 8, 688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Irobi, J.; Almeida-Souza, L.; Asselbergh, B.; de Winter, V.; Goethals, S.; Dierick, I.; Krishnan, J.; Timmermans, J.P.; Robberecht, W.; de Jonghe, P.; et al. Mutant HSPB8 causes motor neuron-specific neurite degeneration. Hum. Mol. Genet. 2010, 19, 3254–3265. [Google Scholar] [CrossRef] [PubMed]
  80. Ackerley, S.; James, P.A.; Kalli, A.; French, S.; Davies, K.E.; Talbot, K. A mutation in the small heat-shock protein HSPB1 leading to distal hereditary motor neuronopathy disrupts neurofilament assembly and the axonal transport of specific cellular cargoes. Hum. Mol. Genet. 2006, 15, 347–354. [Google Scholar] [CrossRef] [Green Version]
  81. Zhai, J.; Lin, H.; Julien, J.P.; Schlaepfer, W.W. Disruption of neurofilament network with aggregation of light neurofilament protein: A common pathway leading to motor neuron degeneration due to Charcot-Marie-Tooth disease-linked mutations in NFL and HSPB. Hum. Mol. Genet. 2007, 16, 3103–3116. [Google Scholar] [CrossRef]
  82. Almeida-Souza, L.; Asselbergh, B.; d’Ydewalle, C.; Moonens, K.; Goethals, S.; de Winter, V.; Azmi, A.; Irobi, J.; Timmermans, J.P.; Gevaert, K.; et al. Small heat-shock protein HSPB1 mutants stabilize microtubules in Charcot-Marie-Tooth neuropathy. J. Neurosci. 2011, 31, 15320–15328. [Google Scholar] [CrossRef] [Green Version]
  83. Kalmar, B.; Innes, A.; Wanisch, K.; Kolaszynska, A.K.; Pandraud, A.; Kelly, G.; Abramov, A.Y.; Reilly, M.M.; Schiavo, G.; Greensmith, L. Mitochondrial deficits and abnormal mitochondrial retrograde axonal transport play a role in the pathogenesis of mutant Hsp27-induced charcot marie tooth disease. Hum. Mol. Genet. 2017, 26, 3313–3326. [Google Scholar] [CrossRef] [Green Version]
  84. Tanabe, H.; Higuchi, Y.; Yuan, J.H.; Hashiguchi, A.; Yoshimura, A.; Ishihara, S.; Nozuma, S.; Okamoto, Y.; Matsuura, E.; Ishiura, H.; et al. Clinical and genetic features of Charcot-Marie-Tooth disease 2F and hereditary motor neuropathy 2B in Japan. J. Peripher. Nerv. Syst. 2018, 23, 40–48. [Google Scholar] [CrossRef] [Green Version]
  85. Lee, J.; Jung, S.-C.; Joo, J.; Choi, Y.-R.; Moon, H.W.; Kwak, G.; Yeo, H.K.; Lee, J.-S.; Ahn, H.-J.; Jung, N.; et al. Overexpression of mutant HSP27 causes axonal neuropathy in mice. J. Biomed. Sci. 2015, 22, 43. [Google Scholar] [CrossRef] [Green Version]
  86. Srivastava, A.K.; Renusch, S.R.; Naiman, N.E.; Gu, S.; Sneh, A.; Arnold, W.D.; Sahenk, Z.; Kolb, S.J. Mutant HSPB1 overexpression in neurons is sufficient to cause age-related motor neuronopathy in mice. Neurobiol. Dis. 2012, 47, 163–173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Muchowski, P.J.; Schaffar, G.; Sittler, A.; Wanker, E.E.; Hayer-Hartl, M.K.; Hartl, F.U. Hsp70 and Hsp40 chaperones can inhibit self-assembly of polyglutamine proteins into amyloid-like fibrils. Proc. Natl. Acad. Sci. USA 2000, 97, 7841–7846. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Jana, N.R.; Tanaka, M.; Wang, G.H.; Nukina, N. Polyglutamine length-dependent interaction of Hsp40 and Hsp70 family chaperones with truncated N-terminal huntingtin: Their role in suppression of aggregation and cellular toxicity. Hum. Mol. Genet. 2000, 9, 2009–2018. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Howarth, J.L.; Kelly, S.; Keasey, M.P.; Glover, C.P.J.; Lee, Y.B.; Mitrophanous, K.; Chapple, J.P.; Gallo, J.M.; Cheetham, M.E.; Uney, J.B. Hsp40 molecules that target to the ubiquitin-proteasome system decrease inclusion formation in models of polyglutamine disease. Mol. Ther. 2007, 15, 1100–1105. [Google Scholar] [CrossRef] [PubMed]
  90. Frasquet, M.; Chumillas, M.J.; Vílchez, J.J.; Márquez-Infante, C.; Palau, F.; Vázquez-Costa, J.F.; Lupo, V.; Espinós, C.; Sevilla, T. Phenotype and natural history of inherited neuropathies caused by HSJ1 c.352+1G>A mutation. J. Neurol. Neurosurg. Psychiatry 2016, 87, 1265–1268. [Google Scholar] [CrossRef]
  91. Greaves, J.; Chamberlain, L.H. Dual role of the cysteine-string domain in membrane binding and palmitoylation-dependent sorting of the molecular chaperone cysteine-string protein. Mol. Biol. Cell 2006, 17, 4748–4759. [Google Scholar] [CrossRef]
  92. Martin, J.J. Adult type of neuronal ceroid-lipofuscinosis. J. Inherit. Metab. Dis. 1993, 16, 237–240. [Google Scholar] [CrossRef]
  93. Haltia, M. The neuronal ceroid-lipofuscinoses. J. Neuropathol. Exp. Neurol. 2003, 62, 1–13. [Google Scholar] [CrossRef] [Green Version]
  94. Xu, F.; Proft, J.; Gibbs, S.; Winkfein, B.; Johnson, J.N.; Syed, N.; Braun, J.E.A. Quercetin targets cysteine string protein (CSPα) and impairs synaptic transmission. PLoS ONE 2010, 5, e11045. [Google Scholar] [CrossRef] [Green Version]
  95. Brocchieri, L.; Conway de Macario, E.; Macario, A.J.L. hsp70 genes in the human genome: Conservation and differentiation patterns predict a wide array of overlapping and specialized functions. BMC EVolume Biol. 2008, 8, 19. [Google Scholar] [CrossRef] [Green Version]
  96. Cheng, M.Y.; Hartl, F.U.; Martin, J.; Pollock, R.A.; Kalousek, F.; Neuper, W.; Hallberg, E.M.; Hallberg, R.L.; Horwich, A.L. Mitochondrial heat-shock protein hsp60 is essential for assembly of proteins imported into yeast mitochondria. Nature 1989, 337, 620–625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Bross, P.; Fernandez-Guerra, P. Disease-associated mutations in the HSPD1 gene encoding the large subunit of the mitochondrial HSP60/HSP10 chaperonin complex. Front. Mol. Biosci. 2016, 3, 49. [Google Scholar] [CrossRef] [PubMed]
  98. Christensen, J.H.; Nielsen, M.N.; Hansen, J.; Füchtbauer, A.; Füchtbauer, E.M.; West, M.; Corydon, T.J.; Gregersen, N.; Bross, P. Inactivation of the hereditary spastic paraplegia-associated Hspd1 gene encoding the Hsp60 chaperone results in early embryonic lethality in mice. Cell Stress Chaperones 2010, 15, 851–863. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Fontaine, B.; Davoine, C.S.; Dürr, A.; Paternotte, C.; Feki, I.; Weissenbach, J.; Hazan, J.; Brice, A. A new locus for autosomal dominant pure spastic paraplegia, on chromosome 2q24-q. Am. J. Hum. Genet. 2000, 66, 702–707. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Bross, P.; Naundrup, S.; Hansen, J.; Nielsen, M.N.; Christensen, J.H.; Kruhøffer, M.; Palmfeldt, J.; Corydon, T.J.; Gregersen, N.; Ang, D.; et al. The Hsp60-(p.V98I) mutation associated with hereditary spastic paraplegia SPG13 compromises chaperonin function both in vitro and in vivo. J. Biol. Chem. 2008, 283, 15694–15700. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Parnas, A.; Nadler, M.; Nisemblat, S.; Horovitz, A.; Mandel, H.; Azem, A. The MitCHAP-60 disease is due to entropic destabilization of the human mitochondrial Hsp60 oligomer. J. Biol. Chem. 2009, 284, 28198–28203. [Google Scholar] [CrossRef] [Green Version]
  102. Vitale, A.M.; Conway de Macario, E.; Alessandro, R.; Cappello, F.; Macario, A.J.L.; Marino Gammazza, A. Missense Mutations of Human Hsp60: A Computational Analysis to Unveil Their Pathological Significance. Front. Genet. 2020, 11, 969. [Google Scholar] [CrossRef]
  103. Bross, P.; Magnoni, R.; Sigaard Bie, A. Molecular Chaperone Disorders: Defective Hsp60 in Neurodegeneration. Curr. Top. Med. Chem. 2013, 12, 2491–2503. [Google Scholar] [CrossRef]
  104. Miyamoto, Y.; Eguchi, T.; Kawahara, K.; Hasegawa, N.; Nakamura, K.; Funakoshi-Tago, M.; Tanoue, A.; Tamura, H.; Yamauchi, J. Hypomyelinating leukodystrophy-associated missense mutation in HSPD1 blunts mitochondrial dynamics. Biochem. Biophys. Res. Commun. 2015, 462, 275–281. [Google Scholar] [CrossRef]
  105. Miyamoto, Y.; Megumi, F.T.; Hasegawa, N.; Eguchi, T.; Tanoue, A.; Tamura, H.; Yamauchi, J. Data supporting mitochondrial morphological changes by SPG13-associated HSPD1 mutants. Data Br. 2016, 6, 482–488. [Google Scholar] [CrossRef] [Green Version]
  106. Mukherjee, K.; Conway de Macario, E.; Macario, A.J.L.; Brocchieri, L. Chaperonin genes on the rise: New divergent classes and intense duplication in human and other vertebrate genomes. BMC EVolume Biol. 2010, 10, 64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Min, W.; Angileri, F.; Luo, H.; Lauria, A.; Shanmugasundaram, M.; Almerico, A.M.; Cappello, F.; Conway de Macario, E.; Lednev, I.K.; Macario, A.J.L.; et al. A human CCT5 gene mutation causing distal neuropathy impairs hexadecamer assembly in an archaeal model. Sci. Rep. 2014, 4, 1–9. [Google Scholar] [CrossRef] [Green Version]
  108. Sergeeva, O.A.; Tran, M.T.; Haase-Pettingell, C.; King, J.A. Biochemical characterization of mutants in chaperonin proteins CCT4 and CCT5 associated with hereditary sensory neuropathy. J. Biol. Chem. 2014, 289, 27470–27480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Spigolon, D.; Gallagher, D.T.; Velazquez-Campoy, A.; Bulone, D.; Narang, J.; San Biagio, P.L.; Cappello, F.; Macario, A.J.L.; Conway de Macario, E.; Robb, F.T. Quantitative analysis of the impact of a human pathogenic mutation on the CCT5 chaperonin subunit using a proxy archaeal ortholog. Biochem. Biophys. Rep. 2017, 12, 66–71. [Google Scholar] [CrossRef] [PubMed]
  110. Nachury, M.V.; Loktev, A.V.; Zhang, Q.; Westlake, C.J.; Peränen, J.; Merdes, A.; Slusarski, D.C.; Scheller, R.H.; Bazan, J.F.; Sheffield, V.C.; et al. A Core Complex of BBS Proteins Cooperates with the GTPase Rab8 to Promote Ciliary Membrane Biogenesis. Cell 2007, 129, 1201–1213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Hildebrandt, F.; Benzing, T.; Katsanis, N. Ciliopathies. N. Engl. J. Med. 2011, 364, 1533–1543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Priya, S.; Nampoothiri, S.; Sen, P.; Sripriya, S. Bardet-Biedl syndrome: Genetics, molecular pathophysiology, and disease management. Indian J. Ophthalmol. 2016, 64, 620–627. [Google Scholar] [CrossRef] [PubMed]
  113. M’Hamdi, O.; Ouertani, I.; Chaabouni-Bouhamed, H. Update on the genetics of bardet-biedl syndrome. Mol. Syndromol. 2014, 5, 51–56. [Google Scholar] [CrossRef]
  114. Kim, J.C.; Ou, Y.Y.; Badano, J.L.; Esmail, M.A.; Leitch, C.C.; Fiedrich, E.; Beales, P.L.; Archibald, J.M.; Katsanis, N.; Rattner, J.B.; et al. MKKS/BBS6, a divergent chaperonin-like protein linked to the obesity disorder Bardet-Biedl syndrome, is a novel centrosomal component required for cytokinesis. J. Cell Sci. 2005, 118, 1007–1020. [Google Scholar] [CrossRef] [Green Version]
  115. Muller, J.; Stoetzel, C.; Vincent, M.C.; Leitch, C.C.; Laurier, V.; Danse, J.M.; Hellé, S.; Marion, V.; Bennouna-Greene, V.; Vicaire, S.; et al. Identification of 28 novel mutations in the Bardet-Biedl syndrome genes: The burden of private mutations in an extensively heterogeneous disease. Hum. Genet. 2010, 127, 583–593. [Google Scholar] [CrossRef] [Green Version]
  116. Billingsley, G.; Bin, J.; Fieggen, K.J.; Duncan, J.L.; Gerth, C.; Ogata, K.; Wodak, S.S.; Traboulsi, E.I.; Fishman, G.A.; Paterson, A.; et al. Mutations in chaperonin-like BBS genes are a major contributor to disease development in a multiethnic Bardet-Biedl syndrome patient population. J. Med. Genet. 2010, 47, 453–463. [Google Scholar] [CrossRef] [PubMed]
  117. Seo, S.; Baye, L.M.; Schulz, N.P.; Beck, J.S.; Zhang, Q.; Slusarski, D.C.; Sheffield, V.C. BBS6, BBS10, and BBS12 form a complex with CCT/TRiC family chaperonins and mediate BBSome assembly. Proc. Natl. Acad. Sci. USA 2010, 107, 1488–1493. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Davis, R.E.; Swiderski, R.E.; Rahmouni, K.; Nishimura, D.Y.; Mullins, R.F.; Agassandian, K.; Philp, A.R.; Searby, C.C.; Andrews, M.P.; Thompson, S.; et al. A knockin mouse model of the Bardet-Biedl syndrome 1 M390R mutation has cilia defects, ventriculomegaly, retinopathy, and obesity. Proc. Natl. Acad. Sci. USA 2007, 104, 19422–19427. [Google Scholar] [CrossRef] [Green Version]
  119. Swiderski, R.E.; Agassandian, K.; Ross, J.L.; Bugge, K.; Cassell, M.D.; Yeaman, C. Structural defects in cilia of the choroid plexus, subfornical organ and ventricular ependyma are associated with ventriculomegaly. Fluids Barriers CNS 2012, 9, 22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Badano, J.L.; Kim, J.C.; Hoskins, B.E.; Lewis, R.A.; Ansley, S.J.; Cutler, D.J.; Castellan, C.; Beales, P.L.; Leroux, M.R.; Katsanis, N. Heterozygous mutations in BBS1, BBS2 and BBS6 have a potential epistatic effect on Bardet-Beidl patients with two mutations at a second BBS locus. Hum. Mol. Genet. 2003, 12, 1651–1659. [Google Scholar] [CrossRef] [Green Version]
  121. Stone, D.L.; Slavotinek, A.; Bouffard, G.G.; Banerjee-Basu, S.; Baxevanis, A.D.; Barr, M.; Biesecker, L.G. Mutation of a gene encoding a putative chaperonin causes McKusick- Kaufman syndrome. Nat. Genet. 2000, 25, 79–82. [Google Scholar] [CrossRef] [PubMed]
  122. Schaefer, E.; Durand, M.; Stoetzel, C.; Doray, B.; Viville, B.; Hellé, S.; Danse, J.M.; Hamel, C.; Bitoun, P.; Goldenberg, A.; et al. Molecular diagnosis reveals genetic heterogeneity for the overlapping MKKS and BBS phenotypes. Eur. J. Med. Genet. 2011, 54, 157–160. [Google Scholar] [CrossRef] [Green Version]
  123. Nitika; Porter, C.M.; Truman, A.W.; Truttmann, M.C. Post-translational modifications of Hsp70 family proteins: Expanding the chaperone code. J. Biol. Chem. 2020, 295, 10689–10708. [Google Scholar] [CrossRef]
  124. Engel, P.; Goepfert, A.; Stanger, F.V.; Harms, A.; Schmidt, A.; Schirmer, T.; Dehio, C. Adenylylation control by intra-or intermolecular active-site obstruction in Fic proteins. Nature 2012, 482, 107–110. [Google Scholar] [CrossRef]
  125. Truttmann, M.C.; Zheng, X.; Hanke, L.; Damon, J.R.; Grootveld, M.; Krakowiak, J.; Pincus, D.; Ploegh, H.L. Unrestrained AMPylation targets cytosolic chaperones and activates the heat shock response. Proc. Natl. Acad. Sci. USA 2017, 114, E152–E160. [Google Scholar] [CrossRef] [Green Version]
  126. Kielkowski, P.; Buchsbaum, I.Y.; Kirsch, V.C.; Bach, N.C.; Drukker, M.; Cappello, S.; Sieber, S.A. FICD activity and AMPylation remodelling modulate human neurogenesis. Nat. Commun. 2020, 11, 517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Truttmann, M.C.; Pincus, D.; Ploegh, H.L. Chaperone AMPylation modulates aggregation and toxicity of neurodegenerative disease-associated polypeptides. Proc. Natl. Acad. Sci. USA 2018, 115, E5008–E5017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Backe, S.J.; Sager, R.A.; Woodford, M.R.; Makedon, A.M.; Mollapour, M. Post-translational modifications of Hsp90 and translating the chaperone code. J. Biol. Chem. 2020, 295, 11099–11117. [Google Scholar] [CrossRef] [PubMed]
  129. Franco, M.C.; Ye, Y.; Refakis, C.A.; Feldman, J.L.; Stokes, A.L.; Basso, M.; De Mera, R.M.M.F.; Sparrow, N.A.; Calingasan, N.Y.; Kiaei, M.; et al. Nitration of Hsp90 induces cell death. Proc. Natl. Acad. Sci. USA 2013, 110, E1102–E1111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Koike, M.; Fukushi, J.; Ichinohe, Y.; Higashimae, N.; Fujishiro, M.; Sasaki, C.; Yamaguchi, M.; Uchihara, T.; Yagishita, S.; Ohizumi, H.; et al. Valosin-containing protein (VCP) in novel feedback machinery between abnormal protein accumulation and transcriptional suppression. J. Biol. Chem. 2010, 285, 21736–21749. [Google Scholar] [CrossRef] [Green Version]
  131. Franco, M.C.; Ricart, K.C.; Gonzalez, A.S.; Dennys, C.N.; Nelson, P.A.; Janes, M.S.; Mehl, R.A.; Landar, A.; Estévez, A.G. Nitration of Hsp90 on Tyrosine 33 regulates mitochondrial metabolism. J. Biol. Chem. 2015, 290, 19055–19066. [Google Scholar] [CrossRef] [Green Version]
  132. Verba, K.A.; Agard, D.A. How Hsp90 and Cdc37 Lubricate Kinase Molecular Switches. Trends Biochem. Sci. 2017, 42, 799–811. [Google Scholar] [CrossRef]
  133. Gracia, L.; Lora, G.; Blair, L.J.; Jinwal, U.K. Therapeutic Potential of the Hsp90/Cdc37 Interaction in Neurodegenerative Diseases. Front. Neurosci. 2019, 13, 1263. [Google Scholar] [CrossRef] [Green Version]
  134. Miyata, Y.; Nishida, E. CK2 Controls Multiple Protein Kinases by Phosphorylating a Kinase-Targeting Molecular Chaperone, Cdc37. Mol. Cell. Biol. 2004, 24, 4065–4074. [Google Scholar] [CrossRef] [Green Version]
  135. Perez, D.I.; Gil, C.; Martinez, A. Protein kinases CK1 and CK2 as new targets for neurodegenerative diseases. Med. Res. Rev. 2011, 31, 924–954. [Google Scholar] [CrossRef] [Green Version]
  136. Zhang, S.H.; Liu, J.; Kobayashi, R.; Tonks, N.K. Identification of the cell cycle regulator VCP (p97/CDC48) as a substrate of the band 4.1-related protein-tyrosine phosphatase PTPH. J. Biol. Chem. 1999, 274, 17806–17812. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Ju, J.S.; Fuentealba, R.A.; Miller, S.E.; Jackson, E.; Piwnica-Worms, D.; Baloh, R.H.; Weihl, C.C. Valosin-containing protein (VCP) is required for autophagy and is disrupted in VCP disease. J. Cell Biol. 2009, 187, 875–888. [Google Scholar] [CrossRef] [PubMed]
  138. Torrecilla, I.; Oehler, J.; Ramadan, K. The role of ubiquitin-dependent segregase p97 (VCP or Cdc48) in chromatin dynamics after DNA double strand breaks. Philos. Trans. R. Soc. B Biol. Sci. 2017, 372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Cloutier, P.; Coulombe, B. Regulation of molecular chaperones through post-translational modifications: Decrypting the chaperone code. Biochim. Biophys. Acta Gene Regul. Mech. 2013, 1829, 443–454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Astakhova, L.N.; Zatsepina, O.G.; Funikov, S.Y.; Zelentsova, E.S.; Schostak, N.G.; Orishchenko, K.E.; Evgen’Ev, M.B.; Garbuz, D.G. Activity of heat shock genes’ promoters in thermally contrasting animal species. PLoS ONE 2015, 10, 115536. [Google Scholar] [CrossRef] [Green Version]
  141. Murphy, M.E. The HSP70 family and cancer. Carcinogenesis 2013, 34, 1181–1188. [Google Scholar] [CrossRef] [Green Version]
  142. Sun, Y.; Zheng, J.; Xu, Y.; Zhang, X. Paraquat-induced inflammatory response of microglia through HSP60/TLR4 signaling. Hum. Exp. Toxicol. 2018, 37, 1161–1168. [Google Scholar] [CrossRef]
  143. Nguyen, P.; Kdeiss, B.; Ord, S.; Hess, K.; Oliverio, R.; Nikolaidis, N. Evolution and natural variation of HSPA1A, the major stress inducible gene, in humans. FASEB J. 2018, 31, 604–618. [Google Scholar]
  144. Hess, K.; Oliverio, R.; Nguyen, P.; Le, D.; Ellis, J.; Kdeiss, B.; Ord, S.; Chalkia, D.; Nikolaidis, N. Concurrent action of purifying selection and gene conversion results in extreme conservation of the major stress-inducible Hsp70 genes in mammals. Sci. Rep. 2018, 8, 5082. [Google Scholar] [CrossRef]
  145. Singh, R.; Kølvraa, S.; Bross, P.; Jensen, U.B.; Gregersen, N.; Tan, Q.; Knudsen, C.; Rattan, S.I.S. Reduced heat shock response in human mononuclear cells during aging and its association with polymorphisms in HSP70 genes. Cell Stress Chaperones 2006, 11, 208–215. [Google Scholar] [CrossRef] [Green Version]
  146. He, M.; Guo, H.; Yang, X.; Zhang, X.; Zhou, L.; Cheng, L.; Zeng, H.; Hu, F.B.; Tanguay, R.M.; Wu, T. Functional SNPs in HSPA1A Gene Predict Risk of Coronary Heart Disease. PLoS ONE 2009, 4, e4851. [Google Scholar] [CrossRef]
  147. Konings, A.; Van Laer, L.; Michel, S.; Pawelczyk, M.; Carlsson, P.I.; Bondeson, M.L.; Rajkowska, E.; Dudarewicz, A.; Vandevelde, A.; Fransen, E.; et al. Variations in HSP70 genes associated with noise-induced hearing loss in two independent populations. Eur. J. Hum. Genet. 2009, 17, 329–335. [Google Scholar] [CrossRef]
  148. Maugeri, N.; Radhakrishnan, J.; Knight, J.C. Genetic determinants of HSP70 gene expression following heat shock. Hum. Mol. Genet. 2010, 19, 4939–4947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Oliverio, R.; Nguyen, P.; Kdeiss, B.; Ord, S.; Daniels, A.J.; Nikolaidis, N. Functional characterization of natural variants found on the major stress inducible 70-kDa heat shock gene, HSPA1A, in humans. Biochem. Biophys. Res. Commun. 2018, 506, 799–804. [Google Scholar] [CrossRef] [PubMed]
  150. Schopf, F.H.; Biebl, M.M.; Buchner, J. The HSP90 chaperone machinery. Nat. Rev. Mol. Cell Biol. 2017, 18, 345–360. [Google Scholar] [CrossRef] [PubMed]
  151. Evans, C.G.; Wisén, S.; Gestwicki, J.E. Heat shock proteins 70 and 90 inhibit early stages of amyloid β-(1-42) aggregation in vitro. J. Biol. Chem. 2006, 281, 33182–33191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Luo, W.; Sun, W.; Taldone, T.; Rodina, A.; Chiosis, G. Heat shock protein 90 in neurodegenerative diseases. Mol. Neurodegener. 2010, 5, 24. [Google Scholar] [CrossRef] [Green Version]
  153. Chen, B.; Piel, W.H.; Gui, L.; Bruford, E.; Monteiro, A. The HSP90 family of genes in the human genome: Insights into their divergence and evolution. Genomics 2005, 86, 627–637. [Google Scholar] [CrossRef] [PubMed]
  154. Barazi, H.O.; Zhou, L.; Templeton, N.S.; Krutzsch, H.C.; Roberts, D.D. Identification of heat shock protein 60 as a molecular mediator of alpha 3 beta 1 integrin activation. Cancer Res. 2002, 62, 1541–1548. [Google Scholar]
  155. Gorska, M.; Marino Gammazza, A.; Zmijewski, M.A.; Campanella, C.; Cappello, F.; Wasiewicz, T.; Kuban-Jankowska, A.; Daca, A.; Sielicka, A.; Popowska, U.; et al. Geldanamycin-Induced Osteosarcoma Cell Death Is Associated with Hyperacetylation and Loss of Mitochondrial Pool of Heat Shock Protein 60 (Hsp60). PLoS ONE 2013, 8, e71135. [Google Scholar] [CrossRef] [Green Version]
  156. Lin, C.Y.; Hu, C.T.; Cheng, C.C.; Lee, M.C.; Pan, S.M.; Lin, T.Y.; Wu, W.S. Oxidation of heat shock protein 60 and protein disulfide isomerase activates ERK and migration of human hepatocellular carcinoma HepG. Oncotarget 2016, 7, 11067–11082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Marino Gammazza, A.; Campanella, C.; Barone, R.; Caruso Bavisotto, C.; Gorska, M.; Wozniak, M.; Carini, F.; Cappello, F.; D’Anneo, A.; Lauricella, M.; et al. Doxorubicin anti-tumor mechanisms include Hsp60 post-translational modifications leading to the Hsp60/p53 complex dissociation and instauration of replicative senescence. Cancer Lett. 2017, 385, 75–86. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Kim, H.S.; Kim, E.M.; Lee, J.; Yang, W.H.; Park, T.Y.; Kim, Y.M.; Cho, J.W. Heat shock protein 60 modified with O-linked N-acetylglucosamine is involved in pancreatic β-cell death under hyperglycemic conditions. FEBS Lett. 2006, 580, 2311–2316. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Chattopadhyay, S.; Mukherjee, A.; Patra, U.; Bhowmick, R.; Basak, T.; Sengupta, S.; Chawla-Sarkar, M. Tyrosine phosphorylation modulates mitochondrial chaperonin Hsp60 and delays rotavirus NSP4-mediated apoptotic signaling in host cells. Cell. Microbiol. 2017, 19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Hayoun, D.; Kapp, T.; Edri-Brami, M.; Ventura, T.; Cohen, M.; Avidan, A.; Lichtenstein, R.G. HSP60 is transported through the secretory pathway of 3-MCA-induced fibrosarcoma tumour cells and undergoes N-glycosylation. FEBS J. 2012, 279, 2083–2095. [Google Scholar] [CrossRef] [Green Version]
  161. Baron, B. Role of the Post-Translational Modifications of HSP60 in Disease; Springer: Cham, Switzerland, 2019; Volume 18, pp. 69–94. [Google Scholar]
  162. Waterhouse, A.; Bertoni, M.; Bienert, S.; Studer, G.; Tauriello, G.; Gumienny, R.; Heer, F.T.; De Beer, T.A.P.; Rempfer, C.; Bordoli, L.; et al. SWISS-MODEL: Homology modelling of protein structures and complexes. Nucleic Acids Res. 2018, 46, W296–W303. [Google Scholar] [CrossRef] [Green Version]
  163. Pettersen, E.F.; Goddard, T.D.; Huang, C.C.; Couch, G.S.; Greenblatt, D.M.; Meng, E.C.; Ferrin, T.E. UCSF Chimera—A visualization system for exploratory research and analysis. J. Comput. Chem. 2004, 25, 1605–1612. [Google Scholar] [CrossRef] [Green Version]
  164. Macario, A.J.L.; Conway de Macario, E. Chaperonopathies and chaperonotherapy. FEBS Lett. 2007, 581, 3681–3688. [Google Scholar] [CrossRef] [Green Version]
  165. Macario, A.J.L.; Conway de Macario, E. Molecular mechanisms in chaperonopathies: Clues to understanding the histopathological abnormalities and developing novel therapies. J. Pathol. 2020, 250, 9–18. [Google Scholar] [CrossRef] [Green Version]
  166. Cappello, F.; Marino Gammazza, A.; Palumbo Piccionello, A.; Campanella, C.; Pace, A.; Conway de Macario, E.; Macario, A.J.L. Hsp60 chaperonopathies and chaperonotherapy: Targets and agents. Expert Opin. Ther. Targets 2014, 18, 185–208. [Google Scholar] [CrossRef]
  167. Meng, Q.; Li, B.X.; Xiao, X. Toward developing chemical modulators of Hsp60 as potential therapeutics. Front. Mol. Biosci. 2018, 5, 35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Macario, A.J.L.; Conway de Macario, E. Hidden chaperonopathies: Alerting physicians and pathologists on the possibility that uncharacteristic, baffling clinical features in otherwise known diseases may be due to failure of the chaperoning system. Life Saf. Secur. 2020. [Google Scholar] [CrossRef]
Figure 1. Evolutionary trees of chaperonin-containing TCP-1 (CCT) chaperonins. (Left): Maximum-likelihood tree of human chaperonins, including CCT monomers, MKKS, BBS10, and BBS12 as well as the members of the CCT8L class, CCT8L1 and CCT8L2. Numbers associated with each branch indicate bootstrap support from 100 replicates. Tree rooted by the archaeal thermosome α subunit of Sulfolobus solfataricus (Ss_ThsA). (Right): Bayesian tree of the same sequences. The numbers assigned to each branch indicate posterior probabilities. Tree rooted by the thermosome α subunit of Thermoplasma acidophilum (Ta_ThsA). Scale bars, number of substitutions per position for a unit branch length. Reproduced from Reference [106]
Figure 1. Evolutionary trees of chaperonin-containing TCP-1 (CCT) chaperonins. (Left): Maximum-likelihood tree of human chaperonins, including CCT monomers, MKKS, BBS10, and BBS12 as well as the members of the CCT8L class, CCT8L1 and CCT8L2. Numbers associated with each branch indicate bootstrap support from 100 replicates. Tree rooted by the archaeal thermosome α subunit of Sulfolobus solfataricus (Ss_ThsA). (Right): Bayesian tree of the same sequences. The numbers assigned to each branch indicate posterior probabilities. Tree rooted by the thermosome α subunit of Thermoplasma acidophilum (Ta_ThsA). Scale bars, number of substitutions per position for a unit branch length. Reproduced from Reference [106]
Applsci 11 00898 g001
Figure 2. Molecular chaperones involved in genetic neurochaperonopathies (NCPs). Various molecular chaperones are involved in genetic NCPs, but only some (orange) are associated with cognitive impairment, while others (blue) are not.
Figure 2. Molecular chaperones involved in genetic neurochaperonopathies (NCPs). Various molecular chaperones are involved in genetic NCPs, but only some (orange) are associated with cognitive impairment, while others (blue) are not.
Applsci 11 00898 g002
Figure 3. Examples of Hsp70(DnaK) and Hsp90 (HSPC) molecular chaperones rendered as light-brown ribbon models, showing the amino acids serine, threonine, and tyrosine in red (ball-and-stick). The models were built using SWISS-MODEL, developed by the Computational Structural Biology Group at the SIB Swiss Institute of Bioinformatics at the Biozentrum, University of Basel (https://swissmodel.expasy.org/, [162]), and were visualized with UCSF Chimera, developed by the Resource for Biocomputing, Visualization, and Informatics at the University of California, San Francisco, with support from NIH P41-GM103311 (https://www.cgl.ucsf.edu/chimera/, [163]). The HSPA1 and HSPA8 models were built using the NP_005336.3 and NP_006588.1, respectively, amino acid sequences reported in the NCBI Protein database, and as template the B chain of the crystal structure PDB ID 3C7N in the RCSB Protein Databank. The structures shown cover the HSPA1 and the HSPA8 sequences between amino acids 3 and 547. The HSPC1 model was built using the amino acid sequence NP_001017963.2 reported in the NCBI Protein database, and as template the B chain of the crystal structure PDB ID 5FWM in the RCSB Protein Databank. The structure shown covers the HSPC1 sequence between amino acids 137 and 821.
Figure 3. Examples of Hsp70(DnaK) and Hsp90 (HSPC) molecular chaperones rendered as light-brown ribbon models, showing the amino acids serine, threonine, and tyrosine in red (ball-and-stick). The models were built using SWISS-MODEL, developed by the Computational Structural Biology Group at the SIB Swiss Institute of Bioinformatics at the Biozentrum, University of Basel (https://swissmodel.expasy.org/, [162]), and were visualized with UCSF Chimera, developed by the Resource for Biocomputing, Visualization, and Informatics at the University of California, San Francisco, with support from NIH P41-GM103311 (https://www.cgl.ucsf.edu/chimera/, [163]). The HSPA1 and HSPA8 models were built using the NP_005336.3 and NP_006588.1, respectively, amino acid sequences reported in the NCBI Protein database, and as template the B chain of the crystal structure PDB ID 3C7N in the RCSB Protein Databank. The structures shown cover the HSPA1 and the HSPA8 sequences between amino acids 3 and 547. The HSPC1 model was built using the amino acid sequence NP_001017963.2 reported in the NCBI Protein database, and as template the B chain of the crystal structure PDB ID 5FWM in the RCSB Protein Databank. The structure shown covers the HSPC1 sequence between amino acids 137 and 821.
Applsci 11 00898 g003
Table 1. Examples of genetic neurochaperonopathies.
Table 1. Examples of genetic neurochaperonopathies.
Mol. Chap.MutationDiseaseM. I.Ref.
HSPA9Homozygous c.376C-T transition in exon 4 resulting in p.Arg126Trp substitutionEVEN-plus syndromeAR[26]
Compound heterozygous for a c.383A-G transition in exon 4 resulting in p.Tyr128Cys substitution, and a 2-bp deletion in exon 8 (c.882_883delAG) causing a frameshift and resulting in a premature termination codon at amino acid 296EVEN-plus syndromeAR[26]
HSPB8Heterozygous c.421A-G transition in exon 1 resulting in p.Lys141Glu substitutiondHMN2AADo[27,28]
Heterozygous c.423G-C transversion in exon 1 resulting in p.Lys141Asn substitutiondHMN2AADo[28,29]
c.423G-T transversion in exon 1 resulting in p.Lys141Asn substitutionCMT2LADo[30,31]
HSPB1c.100G-A transition in exon 1 resulting in p.Gly34Arg substitutiondHMN2BSporadic[32]
c.116C-T transition in exon 1 resulting in p.Pro39Leu substitutiondHMN2B/CMT2FADo[32,33]
c.121G-A transition in exon 1 resulting in p.Glu41Lys substitutiondHMN2BADo[32]
c.250G-C transversion in exon 1 resulting in p.Gly84Arg substitutiondHMN2B/CMT2FADo[33,34,35]
Homozygous c.295C-A transversion in exon 1 resulting in p.Leu99Met substitutiondHMN2B/CMT2FAR[33]
c.379C-T transition in exon 1 resulting in p.Arg127Trp substitutiondHMN2B/CMT2FADo[36,37]
c.404C-G transversion in exon 1 resulting in p.Ser135Cys substitutiondHMN2B/CMT2FADo[38,39]
c.404C-T transition in exon 1 resulting in p.Ser135Phe substitutiondHMN2B/CMT2FADo[33,36,40]
c.406C-T transition in exon 2 resulting in p.Arg136Trp substitutionCMT2FADo[36]
c.407G-T transversion in exon 2 resulting in p.Arg136Leu substitutiondHMN2B/CMT2FSporadic[32]
c.418C-G transversion in exon 2 resulting in p.Arg140Gly substitutiondHMN2B/CMT2FADo or sporadic[33]
c.421A-C transversion in exon 2 resulting in p.Lys141Gln substitutiondHMN2BADo[41]
c.452C-T transition in exon 2 resulting in p.Thr151Ile substitutiondHMN2BADo[36]
c.490A-G transition in exon resulting in exon 3 resulting in p.Thr164Ala substitutionCMT2FADo[42]
c.539C-T transition in exon 3 resulting in p.Thr180Ile substitutiondHMN2B/CMT2FADo[32,43]
c.544C-T transition in exon 3 resulting in p.Pro182Ser substitutiondHMN2BADo[44]
c.545C-T transition in exon 3 resulting in p.Pro182Leu substitutiondHMN2BADo[36]
c.562C-T transition in exon 3 resulting in p.Arg188Trp substitutionCMT2FSporadic[32]
HSPE1Heterozygous c.217C-T transition in exon 2 resulting in p.Leu73Phe substitutionUndefined neurologic disorderSporadic[45]
DNAJB2Homozygous c.14A-G transition in exon 1 resulting in p.Tyr5Cys substitutionCMT2F/DSMA5AR[46]
Homozygous G-A transition in intron 4 (c.229+1G-A)dHMN/DSMA5AR recessive[46]
Homozygous G-A transition in the donor splice site of exon 5 (c.352+1G-A)DSMA5AR[47]
DNAJC5Heterozygous c.344T-G transversion in exon 3 resulting in p.Leu115Arg substitutionCLN4BADo[48,49,50,51]
Heterozygous 3-bp deletion in exon 3 (c.346_348del) resulting in p.Leu116delCLN4BADo[48,49,50,51]
HSPD1Homozygous c.86A-G transition in exon 2 resulting in p.Asp29Gly substitutionHLD4 or MitCHAP-60 diseaseAR[52,53]
Heterozygous c.292G-A transition in exon 3 resulting in p.Val98Ile substitutionSPG13ADo[54]
Heterozygous c.1381C-G transversion in exon 10 resulting in p.Gln461Glu substitutionSPG13ADo[55]
CCT5Homozygous c.440A-G transition in exon 4 resulting in a His147Arg substitutionHereditary sensory neuropathy with spastic paraplegiaAR[56,57]
Homozigous c.670C>G transversion in exon 5 resulting in Leu224Val substitutionDemyelinating neuropathy with severe motor disability.AR[58]
BBS6c.110A-G transition in exon 1 resulting in p.Tyr37Cys substitutionBBSAR[59,60]
c.155G-A in exon 1 transition resulting in p.Gly52Asp substitution BBSAR[61]
c.169A-G transition in exon 1 resulting in Thr57Ala substitutionBBSAR[59]
Homozygous 1-bp deletion (c.281del) in exon 2 resulting in a frameshift after amino acid Phe94 (p.Phe94fs), terminating the protein at amino acid 103BBSAR[59,61]
Homozygosity for a complex 2-bp deletion (c.429_430del and c.433_434del) in exon 2 resulting in a frameshift and a premature termination of the protein at amino acid 157 BBSAR[59,61]
Nonsense mutation leading to premature termination (c.442C-T transition in exon 2 resulting in p.Gln148Ter) BBSAR[60]
c.792T-A transversion in exon 3 resulting in a premature termination (p.Tyr264Ter)BBSAR[61]
c.830C-T transition in exon 3 resulting in p.Leu277Pro substitutionBBSAR[59]
c.1496G-C transversion in exon 3 resulting in p.Cys499Ser substitution BBSAR[60]
BBS10c.32T-G transversion in exon 1 resulting in p.Val11Gly substitutionBBSAR[62]
c.101G-C transversion in exon 1 resulting in p.Arg34Pro substitutionBBSAR[63]
1-bp insertion in exon 2 (c.271dupT) leading to premature termination (p.Cys91fsTer95)BBSAR[63,64]
c.273C-G transversion in exon 2 resulting in p.Cys91Trp substitutionBBSAR[64]
4-bp deletion in exon 2 (c.909_912del) resulting in premature termination (p.S303fsTer305)BBSAR[63]
c.931T-G transversion in exon 2 resulting in p.Ser311Ala substitutionBBSAR[62,63]
2-bp deletion in exon 2 (c.1044_1045del) resulting in a frameshift and premature termination (p.Pro350fs)BBSAR[64]
BBS12Homozygous 3-bp deletion in exon 2 (c.337_339del) resulting in p.Val113delBBSAR[65]
Homozygous c.865G-C transversion in exon 3 resulting in p.Ala289Pro substitutionBBSAR[65]
c.1063C-T transition in exon 3 resulting in a nonsense mutation (p.Arg355Ter)BBSAR[65]
2-bp deletion in exon 3 (c.1115_1116del) resulting in frameshift and premature termination of the protein (p.Phe372fsTer373)BBSAR[65]
2-bp deletion in exon 3 (c.1483_1484del) resulting in frameshift and premature termination (p.Glu495fsTer498)BBSAR[65]
Abbreviations: Mol. Chap., molecular chaperones; M.I., mode of inheritance; Ref., reference; ADo, autosomal dominant; AR, autosomal recessive; dHMN2A/2B, distal Hereditary Motor Neuropathy type 2A/2B; CMT2L/2F, Charcot-Marie-Tooth type 2L/2F; DSMA5, Distal Spinal Muscular Atrophy-5, CLN4B, Ceroid Lipofuscinosis Neuronal 4B; SPG13, Spastic Paraplegia 13; HLD4, Hypomyelinating leukodystrophy-4; BBS, Bardet–Biedl syndrome.
Table 2. Effects of PTM on molecular chaperones produced by in vivo and in vitro wild type and disease models.
Table 2. Effects of PTM on molecular chaperones produced by in vivo and in vitro wild type and disease models.
ModelStatusMolecular ChaperoneHuman OrthologPTMEffectRef.
DrosophilaWtBipGrp78 (HSPA5)AMPylationBlindness[123]
Saccha-romyces cerevisiaeWtHsp70HSPAAMPylationDecreased cell growth; protein misfolding; toxic protein aggregation [125]
Caeno-rhabditis elegansAD; PDHSP-1; HSP-3; HSP-4HSPAAMPylationLarge cytoprotective protein aggregates [127]
PC12 cellsWtHsp90HSPCNitrationALS[128,129]
PC12 cellsWtVCPp97Phosphorylation and acetylationNeurite retraction and shrinkage[130]
Abbreviations: PTM, post-translation modification; Ref., reference; Wt, wild type; Hsp or HSP, heat shock protein; AD, Alzheimer’s disease; PD, Parkinson’s disease; ALS, amyotrophic lateral sclerosis; VCP, valosin-containing protein.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Scalia, F.; Vitale, A.M.; Santonocito, R.; Conway de Macario, E.; Macario, A.J.L.; Cappello, F. The Neurochaperonopathies: Anomalies of the Chaperone System with Pathogenic Effects in Neurodegenerative and Neuromuscular Disorders. Appl. Sci. 2021, 11, 898. https://doi.org/10.3390/app11030898

AMA Style

Scalia F, Vitale AM, Santonocito R, Conway de Macario E, Macario AJL, Cappello F. The Neurochaperonopathies: Anomalies of the Chaperone System with Pathogenic Effects in Neurodegenerative and Neuromuscular Disorders. Applied Sciences. 2021; 11(3):898. https://doi.org/10.3390/app11030898

Chicago/Turabian Style

Scalia, Federica, Alessandra Maria Vitale, Radha Santonocito, Everly Conway de Macario, Alberto J. L. Macario, and Francesco Cappello. 2021. "The Neurochaperonopathies: Anomalies of the Chaperone System with Pathogenic Effects in Neurodegenerative and Neuromuscular Disorders" Applied Sciences 11, no. 3: 898. https://doi.org/10.3390/app11030898

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

Article Metrics

Back to TopTop