Next Article in Journal
Transcriptome Analysis Reveals Altered Expression of Genes Involved in Hypoxia, Inflammation and Immune Regulation in Pdcd10-Depleted Mouse Endothelial Cells
Next Article in Special Issue
High Resolution Analysis of DMPK Hypermethylation and Repeat Interruptions in Myotonic Dystrophy Type 1
Previous Article in Journal
Assessment of Bidirectional Relationships between Leisure Sedentary Behaviors and Neuropsychiatric Disorders: A Two-Sample Mendelian Randomization Study
Previous Article in Special Issue
The Skin We Live in: Pigmentation Traits and Tanning Behaviour in British Young Adults, an Observational and Genetically-Informed Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Multisystem Proteinopathy Due to VCP Mutations: A Review of Clinical Heterogeneity and Genetic Diagnosis

1
Hotchkiss Brain Institute, Department of Clinical Neurosciences, Cumming School of Medicine, University of Calgary, Calgary, AB T2N 4N1, Canada
2
Alberta Child Health Research Institute, Department of Medical Genetics, Cumming School of Medicine, University of Calgary, Calgary, AB T2N 4N1, Canada
3
Division of Genetic and Genomic Medicine, Department of Pediatrics, University of California Irvine Medical Center, Orange, CA 92868, USA
4
Oral Ecology Research Group, Faculty of Dental Medicine, Université Laval, Quebec City, QC G1V 0A6, Canada
5
Cure VCP Disease, Inc., Americus, GA 31709, USA
6
Department of Neurology, Washington University School of Medicine, St. Louis, MO 63110, USA
*
Author to whom correspondence should be addressed.
Genes 2022, 13(6), 963; https://doi.org/10.3390/genes13060963
Submission received: 26 April 2022 / Revised: 24 May 2022 / Accepted: 24 May 2022 / Published: 27 May 2022
(This article belongs to the Special Issue Feature Papers in Human Genomics and Genetic Diseases)

Abstract

:
In this work, we review clinical features and genetic diagnosis of diseases caused by mutations in the gene encoding valosin-containing protein (VCP/p97), the functionally diverse AAA-ATPase. VCP is crucial to a multitude of cellular functions including protein quality control, stress granule formation and clearance, and genomic integrity functions, among others. Pathogenic mutations in VCP cause multisystem proteinopathy (VCP-MSP), an autosomal dominant, adult-onset disorder causing dysfunction in several tissue types. It can result in complex neurodegenerative conditions including inclusion body myopathy, frontotemporal dementia, amyotrophic lateral sclerosis, or combinations of these. There is also an association with other neurodegenerative phenotypes such as Alzheimer-type dementia and Parkinsonism. Non-neurological presentations include Paget disease of bone and may also include cardiac dysfunction. We provide a detailed discussion of genotype-phenotype correlations, recommendations for genetic diagnosis, and genetic counselling implications of VCP-MSP.

1. Introduction

Valosin-containing protein (VCP) is encoded by VCP on chromosome 9 and it is implicated in numerous human disease phenotypes with autosomal dominant inheritance. VCP is a highly conserved, ubiquitously and abundantly expressed ATPase. Indeed, VCP may make up as much as 1% of all cytoplasmic protein [1]. VCP mediates ubiquitin-dependent cellular processes through the ubiquitin-proteasome system (UPS) [2], protein quality control [3,4], transcription factor processing [5], membrane fusion [6], cell cycle control [7], and regulation of autophagy [8,9].
Over 50 heterozygous missense mutations in VCP have been identified in patients with multisystem proteinopathy 1 (MSP1), an autosomal dominant, adult-onset progressive disorder, also known as inclusion body myopathy associated with Paget disease of bone and frontotemporal dementia (IBMPFD), or VCP disease (hereafter referred to as VCP-MSP) [10,11]. The major pathological feature seen in VCP-MSP is the presence of ubiquitin-positive protein aggregates in muscle tissues of patients affected by inclusion body myopathy (IBM). Intranuclear TDP-43+ inclusions are demonstrated in patients with frontotemporal dementia (FTD) and/or amyotrophic lateral sclerosis (ALS) [12]. In Paget disease of bone (PDB), abnormally large osteoclasts containing nuclear inclusions resemble the muscle pathology findings [13]. On the whole, the pathological findings of VCP-MSP suggest that abnormal protein clearance is the major molecular mechanism affected by VCP mutations causing human disease. In other forms of MSP (recently reviewed) [11], it has also been suggested that the common molecular mechanism relates to the disruption of two major protein clearance pathways: the UPS and autophagy [14]. However, future studies may identify others among VCP’s diverse molecular mechanisms that are relevant to MSP and other human diseases. The molecular functions of VCP and disease pathogenesis have recently been reviewed [9].
In this review, we provide an overview of the complex clinical presentations of VCP-MSP, with guidance for genetic diagnosis of this complex condition. A special discussion of the complexities surrounding genetic diagnosis and counselling in VCP-MSP is also included.

2. Genetic Differential for VCP-MSP

The majority of genetically diagnosed families with MSP are associated with VCP mutations [15], although several other genes are associated with MSP phenotypes that may be indistinguishable from VCP-MSP. Mutations in the genes encoding heterogeneous nuclear riboproteins are, respectively, associated with MSP2 and MSP3 [16], although these are thought to be extremely rare [15], with very few reported cases of either disease. MSP2 is caused by mutations in HNRNPA2B1, a gene encoding two ribonucleoproteins by alternative splicing (HNRNPA2 and HNRNPB1), which are required for diverse RNA stabilization, splicing, and translation functions, as well as DNA integrity functions [17]. Most of the phenotypes associated with VCP-MSP have been identified in MSP2 patients (myopathy, FTD, ALS, and PDB) [16]. MSP3 is associated with mutations in HNRNPA1, a gene encoding a nuclear riboprotein with diverse RNA functions and telomere maintenance [18]. MSP3 presents with ALS, myopathy, and PDB phenotypes [16], including some atypical phenotypes such as flail arm presentation of ALS [19].
MSP4 is caused by mutations in SQSTM1, a gene encoding the p62/SQSTM1 protein, a selective autophagy receptor with various other roles in the UPS system and apoptosis [20]. MSP4 has been associated with diverse phenotypes including myopathy [21], PDB [22], ALS [23] and FTD [22]. Interestingly, an uncommon polymorphic variant in TIA1 appears to modify the phenotype of SQSTM1 mutations, shifting the penetrance toward myopathy and manifesting as a late-onset, milder phenotype resembling Welander distal myopathy [24,25]. TIA1 is an RNA-binding protein critical to the cellular stress response [26].
Other genes have been associated with phenotypes resembling MSP and should be considered in the differential diagnosis. MATR3 encodes Matrin 3, a DNA- and RNA-binding protein [27]. MATR3 mutations are linked to familial myopathy with ALS [28], as well as sporadic ALS disease [29,30]. It was also suggested that some of the familial ALS patients may have had associated dementia, though the clinical presentation was not described [28]; the presence of FTD associated with MATR3 mutations is at best very rare [31] and there is only a single well-described case in the literature [32]. There are no cases reported of MATR3 mutations with PDB. TIA1 is associated with Welander distal myopathy [33], caused by a founder mutation present in Scandinavia [34]. More recently, a mutation in TIA1 was associated with ALS and FTD [35], but there has been uncertainty with this finding [36,37]. OPTN encodes Optineurin, an adaptor protein with numerous protein interactions involved in vesicular trafficking [38]. OPTN mutations are another cause of monogenic ALS [39] that appears to have a relationship to other phenotypes resembling MSP, since variation on OPTN is a risk factor for sporadic PDB [40] and has also been connected to cases of FTD [41].

3. VCP and Genotype-Phenotype Correlations

Mutations in VCP cause a spectrum of phenotypes presenting in adulthood, which may coexist with each other in the same patient. We summarize the different clinical features which may appear as manifestations of VCP mutations in Table 1. Myopathy is the most common feature and is present in approximately 90% of patients [42]. This most frequently manifests as proximal weakness but variations in the phenotype are described. PDB is present in up to 50% of patients [43] and is caused by altered bone turnover, resulting in bone deformities and pathological fractures. FTD develops in 30% of cases, characterized by early onset and rapidly progressive cognitive syndrome with early behavioral and language deficits [42]. Approximately 10% of patients develop ALS [42], in which there is multifocal degeneration of motor neurons in the brain and spinal cord causing progressive weakness. Other rare and atypical phenotypes have also been described including Charcot-Marie-Tooth disease type 2Y [44,45], hereditary spastic paraplegia [46], and variations on the myopathy phenotype resembling facioscapulohumeral muscular dystrophy, limb-girdle muscular dystrophy, scapulohumeral muscular dystrophy, oculopharyngeal muscular dystrophy, and distal myopathy [12,47]. Other neurodegenerative conditions may be more common in patients with VCP mutations, including Parkinson disease and Alzheimer disease [42]. Cardiomyopathy is described as an important feature in case series [48] but the prevalence of cardiac dysfunction is unknown. Expert opinion also cites respiratory muscle weakness as an important consideration for screening and management [49,50]. This broad clinical heterogeneity means that VCP-MSP should be considered in numerous different clinical presentations.
Genotype-phenotype correlation remains a major challenge. There have been major efforts to compile clinical and genetic information from large cohorts of patients, but one of the major limitations is that nearly three-quarters of reported cases are caused by mutations at a single codon for the arginine residue at position 155 [42]. So, although there is much variability for these mutations at position 155, with larger numbers, it may become evident that there is a clear phenotypic correlation for mutations at other residues. Thus far, clear genotype-phenotype correlations appear to be limited. The mutations associated with Charcot-Marie-Tooth disease type 2Y appear to cause restricted phenotypes that differ substantially from classical multisystem involvement of multisystem proteinopathy [44,45,51]. Extended follow-up of these patients may still identify some of the classical VCP-MSP manifestations over time. Two mutations at position 395 also appear to be strongly associated with dementia as a single system disorder [52,53], although again this is based on small case numbers and extended follow-up may identify other associations. We have provided a summary of previously reported pathogenic variants in VCP with their genotype-phenotype-mechanistic correlations, organized by their VCP domain in Table 2. These different mutations are also pictorially represented showing their positions in the postulated three-dimensional conformation of the VCP hexamer in Figure 1.
Table 2. Mutations in VCP organized by domain, and genotype-phenotype correlation.
Table 2. Mutations in VCP organized by domain, and genotype-phenotype correlation.
AA SubstitutionNucleotide MutationDomainPhenotype or DiseaseInheritanceReference
D6Vc.17A>TN-Domain
Cofactor and Substrate Binding
Myalgia, myopathy, arrhythmiaAD[54]
I27Vc.79A>GIBM, PDB, FTDAD[55,56]
K60Rc.179A>GALS/FTDSP[57]
D74Vc.221A>TChorea; FTDAD[58]
V87Fc.259G>TIBMSP[59]
R89Wc.265C>TIBM, FTDSP[60]
R89Qc.266G>AALSSP[61]
N91Yc.271A>TIBM, ALS, FTDAD[62]
R93Cc.277C>TIBM, PDB, FTDAD[63]
R93Hc.278G>AHSPSP[64]
R95Cc.283C>TIBMSP[65]
R95Gc.283C>GIBM, PDB, FTD, ALSAD[10,42]
R95Hc.284G>AFTDUN[66]
G97Ec.290G>AIBM, PDB, FTDAD[51,67]
D98Vc.293A>TALS, FTDSP[68]
V99Dc.296T>APDB, FTDAD[69]
I114Vc.340A>GIBM (distal arm), ALSSP[70,71]
P118Lc.353C>TFTDAD[69]
G125Dc.374G>AIBM, PDB, FTDAD[69,72]
I126Fc.376A>TIBMAD[73]
I126Vc.376A>GIBM, FTDSP[59]
T127Ac.379A>GFTDSP[74]
G128Ac.383G>CIBM, PDB, FTDAD[42,75]
G128Cc.382G>TIBMSP[75]
G128Vc.383G>TIBM, PDB, FTDAD[76]
P137Lc.410C>TIBM (distal), PDB, FTDAD[77]
I151Vc.451A>GALSSP[78]
V154Fc.460G>TFTDAD[79]
R155Hc.464G>AIBM, PDB, FTD, ALSAD[10]
R155Cc.463C>TIBM, PDB, FTD, ALS, PDAD[10,80]
R155Pc.464G>CIBM, PDB, FTDAD[10]
R155Sc.463C>AIBM, PDB, FTDAD[81]
R155Lc.464G>TIBM, PDB, FTD, ALS, SNHAD[82]
G156Sc.466G>AIBM, PDB, FTDAD[83,84]
G156Cc.466G>TALSAD[85]
G157Rc.469G>CIBM, PDB, FTD, SNHAD[86]
M158Vc.472A>GPDB, ALS, FTDAD[87,88]
M158Ic.474G>AIBM, PDB, ALSAD[42,75]
R159Hc.476G>AIBM, PDB, FTD, ALSAD
AR
[70,72,89,90]
R159Cc.475C>TIBM, PDB, FTD, ALS, Parkinson disease, HSPAD[80,91,92,93]
R159Gc.475C>GFTD, ALSAD[94]
R159Sc.475C>AFTDAD[88]
A160Pc.478G>CIBM, PDB, FTD, ALSAD[42,75]
E185Kc.553G>ACMT2 [44]
R191Qc.572G>AN-D1 LinkerIBM, PDB, FTD, ALSAD[10]
R191Gc.571C>GIBM, ALSAD[80]
R191Pc.572G>CALSAD[95]
L198Wc.593T>GIBM, PDB, FTDAD[96]
G202Wc.604G>TIBM, FTDAD[97]
I206Fc.616A>TIBM, PDB, FTDAD[98]
A232Ec.695C>AD1
Oligomerization Domain
Heat Enhanced ATPase Domain
IBM, PDB, ALSAD[10]
T262Ac.784A>GIBM, PDB, FTD, ParkinsonismAD[99]
T262Sc.785C>GFTDAD[88]
K386Ec.1156A>GMyopathyNR[100]
N387Hc.1159A>CIBM, FTDAD[96]
N387Tc.1160A>CALSSP[101]
N387Sc.1160G>AIBM, PDB, FTDAD[102]
D395Gc.1184A>GFTDAD[53,79]
D395Ac.1184A>CFTDAD[52]
N401Sc.1202A>GAlzheimer DementiaSP[74]
A439Sc.1315G>TIBM, PDBAD[81]
A439Pc.1315G>CIBM, PDB, FTDAD[103]
A439Gc.1316C>GIBM, FTDAD[97]
R487Hc.1460G>AD2
Major ATPase Domain
FTD, ALSAD[104]
E578Q IBMUN[105]
D592Nc.1774G>AALSAD[94]
R662Cc.1984C>TALSSP[101]
N750Sc.2249A>GALSSP[106]
9:35060456
del (5)
Proposed fs at AA 515
AutismDN[107]
Abbreviations: Amino acid sequence (AA), Autosomal dominant (AD), amyotrophic lateral sclerosis (ALS), Autosomal recessive (AR), Charcot-Marie-Tooth disease type 2 (CMT2), de novo (DN), frontotemporal dementia (FTD), inclusion body myopathy (IBM), Paget disease of bone (PDB), sensorineural hearing loss (SNH), Sporadic (SP), Unknown (UN).
Figure 1. Pathogenic mutations represented on the 3D model of VCP. (A) Three-dimensional representation of the VCP hexamer. We have color-coded individual domains from a single VCP monomer for ease of reference in subsequent images: Green for the N-terminal domain, blue for the N-D1 linker, yellow for the D1 domain, purple for the D1-D2 linker, light blue for the D2 domain, yellow for the C-terminal domain, with grey lines representing the other 5 structural oligomers. In red, we have represented the locations of pathogenic variants in VCP, showing the positions in the N-terminal domain (B), linker domain (C), D1 domain (D), and D2 domain (E). This figure was generated using open source material available from Swiss model [108] and PyMOL [109] based on the proposed structure of the VCP hexamer [110].
Figure 1. Pathogenic mutations represented on the 3D model of VCP. (A) Three-dimensional representation of the VCP hexamer. We have color-coded individual domains from a single VCP monomer for ease of reference in subsequent images: Green for the N-terminal domain, blue for the N-D1 linker, yellow for the D1 domain, purple for the D1-D2 linker, light blue for the D2 domain, yellow for the C-terminal domain, with grey lines representing the other 5 structural oligomers. In red, we have represented the locations of pathogenic variants in VCP, showing the positions in the N-terminal domain (B), linker domain (C), D1 domain (D), and D2 domain (E). This figure was generated using open source material available from Swiss model [108] and PyMOL [109] based on the proposed structure of the VCP hexamer [110].
Genes 13 00963 g001
Reasons for the generally poor genotype-phenotype correlation could be related to various factors. For any late-onset genetic disorder, the interaction of environmental factors is certainly possible, and there is precedent from other late-onset neuromuscular conditions [111,112,113]. Host-microbiome interactions may be another way in which the onset and severity of disease can be modified, as is seen from early work in other neurodegenerative conditions [114]. However, it also bears mention that much of the variability may still be genetic: there is some evidence that modifier effects can alter the phenotype, such as APOε4 alleles associated with the FTD phenotype [115]. A recent publication also suggests a possible interaction with an intermediate-length repeat in ATXN2 and a complex neurodegenerative phenotype [116]. Future studies may also consider other possible modifiers that have been identified from studies of animal models or related disorders [24,25,117].

4. Genetic Diagnosis of VCP-MSP

The diagnosis of VCP-MSP is confirmed by the identification of a pathogenic mutation in VCP using genetic testing. Diagnosis of VCP-MSP should be considered in patients having one or more of the clinical presentations described in Table 1. In the presence of an autosomal dominant family history, the likelihood of diagnosing a genetic disorder is much greater. However, cases without family history may still be caused by VCP mutations due to variations in disease penetrance and expressivity, attribution of alternate diagnoses, or presence of de novo mutations. When interviewing patients for family history, it is important to consider the above-mentioned clinical variability and ask about muscle weakness, bone disease, dementia and ALS, since phenotypes may vary and do not correlate with genotype. As mentioned above, the exception is Charcot-Marie-Tooth disease type 2Y, which appears to be correlated to particular VCP mutations [44,45,51]. Parkinson disease and Alzheimer disease have been associated with VCP mutations, although it is important to be aware that these conditions are also common in the aged population. Some specific phenotypes may have higher likelihood for a diagnosis with VCP-MSP, for example, ALS is a relatively common disease compared with the prevalence of VCP-MSP; therefore, the likelihood of identifying a VCP mutation in a sporadic case of ALS is low. Myopathy with onset in adulthood is extensively heterogeneous with a low likelihood of obtaining any genetic diagnosis; therefore, isolated myopathy on its own would not suggest a high likelihood of identifying VCP-MSP. However, myopathy is the most common presentation of VCP-MSP and it is possible that some cases of VCP-MSP are misdiagnosed as sporadic inclusion body myositis based on reactivity to NT5C1A antibodies [118]. Early-onset FTD is very rare and may portend a higher likelihood of identifying a causative gene mutation.

4.1. Genetic Testing Methods

There are several testing strategies generally available to clinicians, which broadly speaking can be divided into three categories: (a) targeted single gene testing, (b) multi-gene panel sequencing, and (c) unbiased sequencing (exome or whole genome).
Single gene/variant testing: This is when only VCP is sequenced, or if specific variants in VCP are sequenced as a targeted investigation. Generally, this type of approach is only used for testing of family members in which a diagnosis of VCP-MSP is already established in the family. We do not suggest single gene testing in situations where there is not a known diagnosis of VCP-MSP in the family; this is due to heterogeneity of clinical syndromes associated with VCP-MSP that includes a vast genetic differential diagnosis. Therefore, it is more cost- and time-effective to use parallel sequencing approaches for diagnosis of undifferentiated cases. This also applies to cases in which a diagnosis of VCP-MSP is highly likely (e.g., dominant history with PDB, FTD, and myopathy) because there remains a differential diagnosis with other multisystem proteinopathy genes (as per Section 2 above).
Panel sequencing: This is when multiple genes are tested simultaneously, usually applying short-read parallel sequencing methods. Panels are offered by numerous genetic testing companies with clinical certification. Since VCP mutations are associated with so many phenotypes, it is included in panels to investigate genetic causes of myopathy, dementia, and ALS. However, the clinician should be aware that VCP may not be included in panels from all companies. Some panels for skeletal abnormalities, spastic paraplegia, and neuropathies do not routinely include VCP. Since the content of sequencing panels is adjusted over time, panels will hopefully be more inclusive in the future; thus, in patients without an established genetic diagnosis, re-evaluating that VCP was indeed on the initial panel may be warranted, leading to repeat panel testing. In general, for undifferentiated cases, panel sequencing will usually be the most readily available test and the most suitable investigation for diagnosing patients with VCP-MSP or other genetic conditions in the differential diagnosis.
Unbiased sequencing: Genomic approaches that sequence all coding regions (exome sequencing) or the entire genome are typically reserved for use by medical genetic specialists or in research settings. As long as the read depth and coverage are sufficient, VCP will always be sequenced in cases receiving this type of testing. However, because all other genes are also sequenced, this can result in challenges in the interpretation if multiple potentially relevant variants are identified. This also has a high likelihood of identifying incidental genetic findings that have special genetic counselling implications. There is evidence indicating a higher diagnostic yield for unbiased sequencing approaches compared with panel sequencing for genetic diseases, and this may also be the case for VCP-related disorders which can produce unexpected phenotypes.
Our recommendation is that in most situations, multi-gene panel sequencing will be the best option for testing of undifferentiated cases, whereas single gene testing should be reserved for cases with a known family history of VCP-MSP. Broad genetic testing (such as panel or exome/genome) also has the advantage of being able to identify dual (or multiple) genetic diagnoses, which are believed to occur in ~5% of patients with Mendelian genetic disorders [119]. We may eventually become aware of non-Mendelian gene modifier effects or interactions [120], for which broader sequencing approaches will also be necessary.
We recommend that VCP should be included in genetic sequencing panels for a broad range of phenotypes, including myopathy, skeletal disorders, dementia, ALS, spastic paraplegia, and neuropathy, to aid in the identification of a definitive diagnosis for as many patients as possible.

4.2. Access to Genetic Testing

Genetic testing is increasingly accessible and convenient. Many test providers give the option of providing a blood sample or a sample collected at home (saliva or buccal DNA kits). Coverage for the costs of genetic testing differs widely across jurisdictions; however, many companies will offer a reduced price for “patient pay” testing, when insurers or health systems do not cover costs. There are also sponsored genetic testing programs that offer testing free of charge, with several that are currently active for testing of patients with myopathy and neuropathy.

4.3. Complementary Testing and Reverse Clinical Correlation

Our recommendation is to use genetic testing as the means to reach definitive diagnosis. Additional tests such as muscle biopsy, creatine kinase enzyme levels, neurophysiology (nerve conduction testing and electromyography), and skeletal surveys can give supportive information for the diagnosis of VCP-MSP. However, none of these investigations are specific to VCP mutations and, in many cases, typical abnormalities may not be present even in genetically confirmed cases of VCP-MSP [121].
Muscle biopsy findings associated with VCP-MSP classically show rimmed vacuoles, with ubiquitin positive and tubulofilamentous inclusions on electron microscopy [121]. However, these characteristic findings were seen in only 39% of cases from the largest series of muscle biopsies in VCP-MSP [12]. Nonspecific myopathic changes are frequently identified.
MRI of the muscles is increasingly used to differentiate underlying causes of myopathy and aid in genetic differential diagnosis [122]. There are a few previously reported cases with muscle MRI findings in VCP-MSP [122,123,124] and further study will help to better define the range of findings as well as determining their sensitivity and specificity.
These tests may be carried out to investigate patient symptoms, but increasingly these may also be performed as screening investigations or as “reverse clinical correlation” in patients who have a VCP mutation identified in genetic testing. The former situation is familiar and straightforward, where an undifferentiated patient with myopathy symptoms and signs on examination is referred for a muscle biopsy, or a patient with muscle atrophy and fasciculations is referred for electromyography. When additional testing is requested in a patient already diagnosed with genetic testing, this is often variably based on clinician experience and individual judgement. For example, a patient with myopathy who is found to have a pathogenic VCP mutation on a multi-gene panel may be referred for electromyography, skeletal survey, and MRI of the head, to identify other features in the clinical spectrum of VCP-MSP prior to the development of symptoms. This can have particular value for early identification of PDB, which is responsive to therapy with bisphosphonates [125]. Identification of other features such as ALS may have value for prognostication, or as a baseline investigation that can be a useful reference in case of later clinical deterioration.

4.4. Interpreting Genetic Test Results

The general principles of interpreting genetic variants in VCP are similar to those for other diseases. Most commonly, the American College of Medical Genetics (ACMG) criteria [126] are applied to reach a determination.
(1) 
Positive: A positive result means that a variant in VCP known to be associated with VCP-MSP has been identified in the patient (i.e., a “pathogenic” or “likely pathogenic” variant). Thus far, all confirmed pathogenic variants in VCP are missense. A positive result may confirm the diagnosis in a person who meets clinical criteria or provide a pre-symptomatic diagnosis for an unaffected individual. If an unaffected individual has a positive result, there is a 90% chance that he or she will experience one or more of the features of VCP-MSP by the age of 45 [121]. Onset of symptoms is age-dependent and pre-symptomatic patients should be monitored by a healthcare professional on a regular basis.
(2) 
Negative: A negative result means that no pathogenic variant has been identified in VCP. Depending on the test that was performed, and the purpose of testing, the implications of this result may vary:
(a) 
Affected Individual: This means their condition is not caused by a mutation in VCP, but it remains possible that a mutation in another gene that was not tested could be responsible.
(b) 
Unaffected individual in a family with a known pathogenic VCP mutation: In this situation, a negative result means that the person does not have the mutation and, therefore, will not develop VCP-MSP. The individual will not require follow-up care or monitoring for symptoms as they are confirmed to be unaffected and cannot pass it on to their children.
(c) 
Affected individual in a family having a pathogenic VCP mutation: Individuals who are symptomatic but are non-carriers of the familial mutation may also have a “phenocopy” syndrome that should be considered in this situation [127]. Causes of phenocopy syndromes could be due to diagnostic confusion with acquired disease, a different genetic diagnosis, or non-organic symptoms.

4.5. Variant of Unknown Significance

A variant of unknown significance (VUS) is a change in VCP that has certain characteristics of a pathogenic mutation (e.g., not present in controls and alters the amino acid sequence) but has not previously been associated with human disease. This result means that insufficient data exist to support whether the variant is benign or pathogenic. If it has been identified previously, the variant may have conflicting reports of pathogenicity. The lab evaluates the variant to determine the consequence of the amino acid sequence, level of conservation and frequency in population databases, and will provide this information in the report. Because of the uncertainty of their relationship to disease, VUS are not clinically actionable. VUS can sometimes be reclassified based on results from familial segregation (when sufficient numbers of affected/unaffected individuals are available) or research-based investigations. The lab may also reclassify a VUS after new information becomes available; many labs will issue updated reports when this occurs, but it can also be appropriate to recontact the lab time for VUS reanalysis. In general, genetic testing reports will attempt to integrate all available current information to determine a variant’s classification, but it is worth mentioning that information on public databases such as ClinVar or HGMD may not necessarily be up to date. The ACMG criteria also contain additional discussion of VUS for further reference [126].

5. Genetic Counselling

For undifferentiated cases, ordering physicians should ensure patients are aware that genetic test results may have implications for other family members and that depending on results, formal genetic counselling may be recommended for themselves as well as other family members. In these respects (and relating to other issues discussed below), genetic testing has been considered to have broader implications than other investigations (“genetic exceptionalism”), although in the context of wide availability of unbiased or parallel sequencing tests, this topic has also been subject to evolving debate in the literature [128,129]. Given the high likelihood of identifying VUS with panel or exome/genome sequencing, patients should also be prepared for the possibility of an indeterminate result.
Access to formal genetic counselling in some regions has not grown at the same rate as access to genetic testing [130]. Various approaches have been recommended including the use of telehealth to improve access to these resources [131]. There are also online resources available that can provide genetic counselling, and evidence shows this approach can be highly satisfactory for patients [132].

5.1. Risk to Other Family Members

Each cell in the body has two copies of the VCP gene. VCP-MSP is an autosomal dominant disorder, meaning that one pathogenic mutation in either copy of VCP is sufficient to cause the disease. If a parent has a pathogenic VCP variant, then each of his or her children will have a 50% probability (1 in 2 chance) of inheriting the variant, regardless of sex. If a patient has a pathogenic VCP variant, each full sibling has a 50% probability (1 in 2 chance) of carrying the variant as well, unless it is a confirmed de novo variant.
The known VCP pathogenic variants are highly penetrant, with approximately 90% of mutation carriers developing symptoms by the age of 45 [121]. Variability in severity and age of onset is seen, even among members of the same family with the same variant. For example, the disease may progress at different rates, and family members may have variable manifestations such as PDB, FTD, and/or ALS [62]. This variability is not currently understood but may be related to genetic modifiers or environmental factors.

5.2. Pre-Symptomatic Testing

If there is a known VCP variant in a family with VCP-MSP, testing is available for pre-symptomatic individuals. Individuals may wish to pursue pre-symptomatic testing to resolve uncertainty regarding their risk of the disease, including reproductive, career, or other lifestyle planning [133]. In general, such testing should be a carefully considered decision following formal pre-test genetic counselling and thorough review of the implications of a positive or negative result [134]. Post-test genetic counselling is equally important for individuals who test positive or negative with pre-symptomatic testing [135]. Offering mental health supports can be highly beneficial, again irrespective of the test result [135].
One study specifically considered pre-symptomatic testing for individuals from families with known VCP-MSP. Anxiety levels were initially high before and after testing, but back to their normal anxiety level one year following testing [136]. Predictive testing for pre-symptomatic individuals is an emotionally complex process for the patient, and testing protocols for other conditions such as Huntington disease may be referenced and adapted for VCP-MSP [137]. An unaffected individual that tests negative while other family members test positive may feel guilt for their unaffected status.
The presence of a pathogenic VCP variant in an asymptomatic individual identifies them as a pre-symptomatic patient. Because of the above-mentioned high penetrance for VCP mutations, pre-symptomatic individuals should be monitored for the development of disease manifestations.
The diagnosis of VCP-MSP poses some unique challenges compared to other genetic disorders. The most common clinical presentation will be myopathy, which is generally slowly progressive. In the course of discussing a VCP-MSP diagnosis, it is also necessary to mention the chance of eventually developing a more severe disorder such as frontotemporal dementia or ALS. This effectively results in a situation where a symptomatic patient with VCP-MSP also simultaneously becomes a pre-symptomatic patient for a separate, severe, variable-penetrance disease. There may be additional unexpected psychosocial consequences to patients in this situation, for example, the myopathy patient who repeatedly becomes concerned that they have started developing ALS if they notice any hand or bulbar symptoms at all or likewise are repeatedly concerned that they have started to develop FTD if they perceive anything that could be a cognitive symptom. We emphasize the value of continued follow-up with patients and proactive engagement of mental health resources.

5.3. Genetic Discrimination

Fear of genetic discrimination and protection of personal data are major factors present among individuals considering genetic testing [138]. Furthermore, a person’s culture, race, and ethnicity may influence their willingness to pursue genetic testing, so it may be important to address genetic discrimination fears and cultural influences during genetic counselling [139].
Genetic discrimination and the laws to protect patients vary widely depending on a patient’s country, and many countries have started to adopt laws to protect patients [140]. Canada’s Genetic Non-Discrimination Act (GNDA) prohibits companies and employers from requiring genetic testing or the results of genetic tests. In Europe, a collection of guidelines have been adopted concerning the use of genetic information for insurance purposes [141]. In the United States, all individuals pursuing genetic testing are protected against genetic discrimination by the Genetic Information Nondiscrimination Act of 2008 (GINA) [142]. This act protects individuals from discrimination based on their genetic status with employment and health insurance. Employers are prevented from using employee genetic information to make employment decisions and are prevented from requesting genetic information from employees. Health insurers are prevented from denying coverage, making decisions on premiums or coverage based on genetic information, or requesting genetic information from individuals or family members. GINA does not include protections from life, disability, or long-term care insurance discrimination. There are exceptions to the employment and health insurance protections explained above, and relevant details should be discussed with patients depending on their specific circumstances and jurisdiction.

6. Conclusions

The complexity of VCP’s diverse molecular functions is also mirrored by the variability in clinical dysfunction caused by pathogenic variants in VCP. The relationship between specific molecular functions of VCP and the spectrum of clinical presentations remains poorly understood and, in general, genotype–phenotype correlation is still difficult to demonstrate. Because of this, the diagnosis of VCP-MSP is best achieved by the inclusion of VCP in numerous sequencing panels including myopathy, PDB, FTD, ALS, axonal Charcot-Marie-Tooth disease, and/or spastic paraplegia. The inclusion of VCP in panels for common neurodegenerative disorders such as Alzheimer and Parkinson disease may also be reasonable, but clinical judgement is required in test result interpretation for these disorders which are common in the aged population. As a late-onset disorder with variable disease severity and no disease-modifying therapy, the implications of genetic testing for patients and family members are complex, and patients benefit from access to genetic counselling to discuss the unique aspects of their test implications.

Author Contributions

G.P. and G.L. authored the manuscript. G.P. contributed to Table 1. C.C.W., C.S.P. and G.P. contributed to Table 2. C.S.P. contributed to Figure 1. C.S.P., R.D.F., A.P., C.C.W. and V.K. revised the article for important intellectual content. A.P. and V.K. conceptualized and supervised the project. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All relevant data have been cited within the document.

Acknowledgments

C.S.P. is the recipient of an Eyes High Doctoral Recruitment Scholarship from the Faculty of Graduate Studies at University of Calgary. We acknowledge Ryan P. Mahoney for critical reading of an earlier version of the article.

Conflicts of Interest

The authors declare that they have no competing interests to report.

References

  1. Dalal, S.; Hanson, P.I. Membrane traffic: What drives the AAA motor? Cell 2001, 104, 5–8. [Google Scholar] [CrossRef] [Green Version]
  2. Meyer, H.; Weihl, C.C. The VCP/p97 system at a glance: Connecting cellular function to disease pathogenesis. J. Cell Sci. 2014, 127, 3877–3883. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Rabinovich, E.; Kerem, A.; Frohlich, K.U.; Diamant, N.; Bar-Nun, S. AAA-ATPase p97/Cdc48p, a cytosolic chaperone required for endoplasmic reticulum-associated protein degradation. Mol. Cell. Biol. 2002, 22, 626–634. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Shih, Y.T.; Hsueh, Y.P. VCP and ATL1 regulate endoplasmic reticulum and protein synthesis for dendritic spine formation. Nat. Commun. 2016, 7, 11020. [Google Scholar] [CrossRef] [Green Version]
  5. Rape, M.; Hoppe, T.; Gorr, I.; Kalocay, M.; Richly, H.; Jentsch, S. Mobilization of processed, membrane-tethered SPT23 transcription factor by CDC48(UFD1/NPL4), a ubiquitin-selective chaperone. Cell 2001, 107, 667–677. [Google Scholar] [CrossRef] [Green Version]
  6. Uchiyama, K.; Kondo, H. p97/p47-Mediated biogenesis of Golgi and ER. J. Biochem. 2005, 137, 115–119. [Google Scholar] [CrossRef]
  7. Frohlich, K.U.; Fries, H.W.; Rudiger, M.; Erdmann, R.; Botstein, D.; Mecke, D. Yeast cell cycle protein CDC48p shows full-length homology to the mammalian protein VCP and is a member of a protein family involved in secretion, peroxisome formation, and gene expression. J. Cell Biol. 1991, 114, 443–453. [Google Scholar] [CrossRef] [Green Version]
  8. 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]
  9. Ferrari, V.; Cristofani, R.; Tedesco, B.; Crippa, V.; Chierichetti, M.; Casarotto, E.; Cozzi, M.; Mina, F.; Piccolella, M.; Galbiati, M.; et al. Valosin Containing Protein (VCP): A Multistep Regulator of Autophagy. Int. J. Mol. Sci. 2022, 23, 1939. [Google Scholar] [CrossRef]
  10. Watts, G.D.; Wymer, J.; Kovach, M.J.; Mehta, S.G.; Mumm, S.; Darvish, D.; Pestronk, A.; Whyte, M.P.; Kimonis, V.E. Inclusion body myopathy associated with Paget disease of bone and frontotemporal dementia is caused by mutant valosin-containing protein. Nat. Genet. 2004, 36, 377–381. [Google Scholar] [CrossRef]
  11. Korb, M.K.; Kimonis, V.E.; Mozaffar, T. Multisystem proteinopathy: Where myopathy and motor neuron disease converge. Muscle Nerve 2021, 63, 442–454. [Google Scholar] [CrossRef] [PubMed]
  12. Kimonis, V.E.; Mehta, S.G.; Fulchiero, E.C.; Thomasova, D.; Pasquali, M.; Boycott, K.; Neilan, E.G.; Kartashov, A.; Forman, M.S.; Tucker, S.; et al. Clinical studies in familial VCP myopathy associated with Paget disease of bone and frontotemporal dementia. Am. J. Med. Genet. A 2008, 146, 745–757. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Kimonis, V.E.; Fulchiero, E.; Vesa, J.; Watts, G. VCP disease associated with myopathy, Paget disease of bone and frontotemporal dementia: Review of a unique disorder. Biochim. Biophys. Acta 2008, 1782, 744–748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Taylor, J.P. Multisystem proteinopathy: Intersecting genetics in muscle, bone, and brain degeneration. Neurology 2015, 85, 658–660. [Google Scholar] [CrossRef]
  15. Le Ber, I.; Van Bortel, I.; Nicolas, G.; Bouya-Ahmed, K.; Camuzat, A.; Wallon, D.; De Septenville, A.; Latouche, M.; Lattante, S.; Kabashi, E.; et al. hnRNPA2B1 and hnRNPA1 mutations are rare in patients with “multisystem proteinopathy” and frontotemporal lobar degeneration phenotypes. Neurobiol. Aging 2014, 35, 934.e5–934.e6. [Google Scholar] [CrossRef]
  16. Kim, H.J.; Kim, N.C.; Wang, Y.D.; Scarborough, E.A.; Moore, J.; Diaz, Z.; MacLea, K.S.; Freibaum, B.; Li, S.; Molliex, A.; et al. Mutations in prion-like domains in hnRNPA2B1 and hnRNPA1 cause multisystem proteinopathy and ALS. Nature 2013, 495, 467–473. [Google Scholar] [CrossRef]
  17. Liu, Y.; Shi, S.L. The roles of hnRNP A2/B1 in RNA biology and disease. Wiley Interdiscip. Rev. RNA 2021, 12, e1612. [Google Scholar] [CrossRef]
  18. Clarke, J.P.; Thibault, P.A.; Salapa, H.E.; Levin, M.C. A Comprehensive Analysis of the Role of hnRNP A1 Function and Dysfunction in the Pathogenesis of Neurodegenerative Disease. Front. Mol. Biosci. 2021, 8, 659610. [Google Scholar] [CrossRef]
  19. Liu, Q.; Shu, S.; Wang, R.R.; Liu, F.; Cui, B.; Guo, X.N.; Lu, C.X.; Li, X.G.; Liu, M.S.; Peng, B.; et al. Whole-exome sequencing identifies a missense mutation in hnRNPA1 in a family with flail arm ALS. Neurology 2016, 87, 1763–1769. [Google Scholar] [CrossRef]
  20. Kumar, A.V.; Mills, J.; Lapierre, L.R. Selective Autophagy Receptor p62/SQSTM1, a Pivotal Player in Stress and Aging. Front. Cell Dev. Biol. 2022, 10, 793328. [Google Scholar] [CrossRef]
  21. Bucelli, R.C.; Arhzaouy, K.; Pestronk, A.; Pittman, S.K.; Rojas, L.; Sue, C.M.; Evila, A.; Hackman, P.; Udd, B.; Harms, M.B.; et al. SQSTM1 splice site mutation in distal myopathy with rimmed vacuoles. Neurology 2015, 85, 665–674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Rea, S.L.; Majcher, V.; Searle, M.S.; Layfield, R. SQSTM1 mutations--bridging Paget disease of bone and ALS/FTLD. Exp. Cell Res. 2014, 325, 27–37. [Google Scholar] [CrossRef] [PubMed]
  23. Fecto, F.; Yan, J.; Vemula, S.P.; Liu, E.; Yang, Y.; Chen, W.; Zheng, J.G.; Shi, Y.; Siddique, N.; Arrat, H.; et al. SQSTM1 mutations in familial and sporadic amyotrophic lateral sclerosis. Arch. Neurol. 2011, 68, 1440–1446. [Google Scholar] [CrossRef] [PubMed]
  24. Lee, Y.; Jonson, P.H.; Sarparanta, J.; Palmio, J.; Sarkar, M.; Vihola, A.; Evila, A.; Suominen, T.; Penttila, S.; Savarese, M.; et al. TIA1 variant drives myodegeneration in multisystem proteinopathy with SQSTM1 mutations. J. Clin. Investig. 2018, 128, 1164–1177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Niu, Z.; Pontifex, C.S.; Berini, S.; Hamilton, L.E.; Naddaf, E.; Wieben, E.; Aleff, R.A.; Martens, K.; Gruber, A.; Engel, A.G.; et al. Myopathy With SQSTM1 and TIA1 Variants: Clinical and Pathological Features. Front. Neurol. 2018, 9, 147. [Google Scholar] [CrossRef] [PubMed]
  26. Gilks, N.; Kedersha, N.; Ayodele, M.; Shen, L.; Stoecklin, G.; Dember, L.M.; Anderson, P. Stress granule assembly is mediated by prion-like aggregation of TIA-1. Mol. Biol. Cell 2004, 15, 5383–5398. [Google Scholar] [CrossRef] [Green Version]
  27. Malik, A.M.; Barmada, S.J. Matrin 3 in neuromuscular disease: Physiology and pathophysiology. JCI Insight 2021, 6, e143948. [Google Scholar] [CrossRef]
  28. Johnson, J.O.; Pioro, E.P.; Boehringer, A.; Chia, R.; Feit, H.; Renton, A.E.; Pliner, H.A.; Abramzon, Y.; Marangi, G.; Winborn, B.J.; et al. Mutations in the Matrin 3 gene cause familial amyotrophic lateral sclerosis. Nat. Neurosci. 2014, 17, 664–666. [Google Scholar] [CrossRef]
  29. Lin, K.P.; Tsai, P.C.; Liao, Y.C.; Chen, W.T.; Tsai, C.P.; Soong, B.W.; Lee, Y.C. Mutational analysis of MATR3 in Taiwanese patients with amyotrophic lateral sclerosis. Neurobiol. Aging 2015, 36, 2005.e1–2005.e4. [Google Scholar] [CrossRef]
  30. Leblond, C.S.; Gan-Or, Z.; Spiegelman, D.; Laurent, S.B.; Szuto, A.; Hodgkinson, A.; Dionne-Laporte, A.; Provencher, P.; de Carvalho, M.; Orru, S.; et al. Replication study of MATR3 in familial and sporadic amyotrophic lateral sclerosis. Neurobiol. Aging 2016, 37, 209.e17–209.e21. [Google Scholar] [CrossRef]
  31. Millecamps, S.; De Septenville, A.; Teyssou, E.; Daniau, M.; Camuzat, A.; Albert, M.; LeGuern, E.; Galimberti, D.; French research network on, F.T.D.; Ftd, A.L.S.; et al. Genetic analysis of matrin 3 gene in French amyotrophic lateral sclerosis patients and frontotemporal lobar degeneration with amyotrophic lateral sclerosis patients. Neurobiol. Aging 2014, 35, 2882.e13–2882.e15. [Google Scholar] [CrossRef] [PubMed]
  32. Castro, M.; Venkateswaran, N.; Peters, S.T.; Deyle, D.R.; Bower, M.; Koob, M.D.; Boeve, B.F.; Vossel, K. Case Report: Early-Onset Behavioral Variant Frontotemporal Dementia in Patient With Retrotransposed Full-Length Transcript of Matrin-3 Variant 5. Front. Neurol. 2020, 11, 600468. [Google Scholar] [CrossRef]
  33. Hackman, P.; Sarparanta, J.; Lehtinen, S.; Vihola, A.; Evila, A.; Jonson, P.H.; Luque, H.; Kere, J.; Screen, M.; Chinnery, P.F.; et al. Welander distal myopathy is caused by a mutation in the RNA-binding protein TIA1. Ann. Neurol. 2013, 73, 500–509. [Google Scholar] [CrossRef] [PubMed]
  34. Klar, J.; Sobol, M.; Melberg, A.; Mabert, K.; Ameur, A.; Johansson, A.C.; Feuk, L.; Entesarian, M.; Orlen, H.; Casar-Borota, O.; et al. Welander distal myopathy caused by an ancient founder mutation in TIA1 associated with perturbed splicing. Hum. Mutat. 2013, 34, 572–577. [Google Scholar] [CrossRef] [PubMed]
  35. Mackenzie, I.R.; Nicholson, A.M.; Sarkar, M.; Messing, J.; Purice, M.D.; Pottier, C.; Annu, K.; Baker, M.; Perkerson, R.B.; Kurti, A.; et al. TIA1 Mutations in Amyotrophic Lateral Sclerosis and Frontotemporal Dementia Promote Phase Separation and Alter Stress Granule Dynamics. Neuron 2017, 95, 808–816.e9. [Google Scholar] [CrossRef] [PubMed]
  36. van der Spek, R.A.; van Rheenen, W.; Pulit, S.L.; Kenna, K.P.; Ticozzi, N.; Kooyman, M.; McLaughlin, R.L.; Moisse, M.; van Eijk, K.R.; van Vugt, J.; et al. Reconsidering the causality of TIA1 mutations in ALS. Amyotroph. Lateral Scler. Frontotemporal. Degener. 2018, 19, 1–3. [Google Scholar] [CrossRef]
  37. Baradaran-Heravi, Y.; Dillen, L.; Nguyen, H.P.; Van Mossevelde, S.; Baets, J.; De Jonghe, P.; Engelborghs, S.; De Deyn, P.P.; Vandenbulcke, M.; Vandenberghe, R.; et al. No supportive evidence for TIA1 gene mutations in a European cohort of ALS-FTD spectrum patients. Neurobiol. Aging 2018, 69, 293.e9–293.e11. [Google Scholar] [CrossRef]
  38. Ryan, T.A.; Tumbarello, D.A. Optineurin: A Coordinator of Membrane-Associated Cargo Trafficking and Autophagy. Front. Immunol. 2018, 9, 1024. [Google Scholar] [CrossRef]
  39. Maruyama, H.; Morino, H.; Ito, H.; Izumi, Y.; Kato, H.; Watanabe, Y.; Kinoshita, Y.; Kamada, M.; Nodera, H.; Suzuki, H.; et al. Mutations of optineurin in amyotrophic lateral sclerosis. Nature 2010, 465, 223–226. [Google Scholar] [CrossRef]
  40. Albagha, O.M.; Visconti, M.R.; Alonso, N.; Langston, A.L.; Cundy, T.; Dargie, R.; Dunlop, M.G.; Fraser, W.D.; Hooper, M.J.; Isaia, G.; et al. Genome-wide association study identifies variants at CSF1, OPTN and TNFRSF11A as genetic risk factors for Paget’s disease of bone. Nat. Genet. 2010, 42, 520–524. [Google Scholar] [CrossRef] [Green Version]
  41. Pottier, C.; Bieniek, K.F.; Finch, N.; van de Vorst, M.; Baker, M.; Perkersen, R.; Brown, P.; Ravenscroft, T.; van Blitterswijk, M.; Nicholson, A.M.; et al. Whole-genome sequencing reveals important role for TBK1 and OPTN mutations in frontotemporal lobar degeneration without motor neuron disease. Acta Neuropathol. 2015, 130, 77–92. [Google Scholar] [CrossRef] [PubMed]
  42. Al-Obeidi, E.; Al-Tahan, S.; Surampalli, A.; Goyal, N.; Wang, A.K.; Hermann, A.; Omizo, M.; Smith, C.; Mozaffar, T.; Kimonis, V. Genotype-phenotype study in patients with valosin-containing protein mutations associated with multisystem proteinopathy. Clin. Genet. 2018, 93, 119–125. [Google Scholar] [CrossRef] [PubMed]
  43. Farpour, F.; Tehranzadeh, J.; Donkervoort, S.; Smith, C.; Martin, B.; Vanjara, P.; Osann, K.; Kimonis, V.E. Radiological features of Paget disease of bone associated with VCP myopathy. Skeletal. Radiol. 2012, 41, 329–337. [Google Scholar] [CrossRef] [PubMed]
  44. Gonzalez, M.A.; Feely, S.M.; Speziani, F.; Strickland, A.V.; Danzi, M.; Bacon, C.; Lee, Y.; Chou, T.F.; Blanton, S.H.; Weihl, C.C.; et al. A novel mutation in VCP causes Charcot-Marie-Tooth Type 2 disease. Brain 2014, 137, 2897–2902. [Google Scholar] [CrossRef] [PubMed]
  45. Gite, J.; Milko, E.; Brady, L.; Baker, S.K. Phenotypic convergence in Charcot-Marie-Tooth 2Y with novel VCP mutation. Neuromuscul. Disord. 2020, 30, 232–235. [Google Scholar] [CrossRef]
  46. van de Warrenburg, B.P.; Schouten, M.I.; de Bot, S.T.; Vermeer, S.; Meijer, R.; Pennings, M.; Gilissen, C.; Willemsen, M.A.; Scheffer, H.; Kamsteeg, E.J. Clinical exome sequencing for cerebellar ataxia and spastic paraplegia uncovers novel gene-disease associations and unanticipated rare disorders. Eur. J. Hum. Genet. 2017, 25, 393. [Google Scholar] [CrossRef] [Green Version]
  47. Kimonis, V.E.; Watts, G.D. Autosomal dominant inclusion body myopathy, Paget disease of bone, and frontotemporal dementia. Alzheimer Dis. Assoc. Disord. 2005, 19 (Suppl. S1), S44–S47. [Google Scholar] [CrossRef]
  48. Miller, T.D.; Jackson, A.P.; Barresi, R.; Smart, C.M.; Eugenicos, M.; Summers, D.; Clegg, S.; Straub, V.; Stone, J. Inclusion body myopathy with Paget disease and frontotemporal dementia (IBMPFD): Clinical features including sphincter disturbance in a large pedigree. J. Neurol. Neurosurg. Psychiatry 2009, 80, 583–584. [Google Scholar] [CrossRef]
  49. Pfeffer, G.; Povitz, M. Respiratory management of patients with neuromuscular disease: Current perspectives. Degener Neurol Neuromuscul. Dis. 2016, 6, 111–118. [Google Scholar] [CrossRef] [Green Version]
  50. Korb, M.; Peck, A.; Alfano, L.N.; Berger, K.I.; James, M.K.; Ghoshal, N.; Healzer, E.; Henchcliffe, C.; Khan, S.; Mammen, P.P.A.; et al. Development of a standard of care for patients with valosin-containing protein associated multisystem proteinopathy. Orphanet J. Rare Dis. 2022, 17, 23. [Google Scholar] [CrossRef]
  51. Jerath, N.U.; Crockett, C.D.; Moore, S.A.; Shy, M.E.; Weihl, C.C.; Chou, T.F.; Grider, T.; Gonzalez, M.A.; Zuchner, S.; Swenson, A. Rare Manifestation of a c.290 C>T, p.Gly97Glu VCP Mutation. Case Rep. Genet. 2015, 2015, 239167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Bruno, F.; Conidi, M.E.; Puccio, G.; Frangipane, F.; Lagana, V.; Bernardi, L.; Smirne, N.; Mirabelli, M.; Colao, R.; Curcio, S.; et al. A Novel Mutation (D395A) in Valosin-Containing Protein Gene Is Associated With Early Onset Frontotemporal Dementia in an Italian Family. Front. Genet. 2021, 12, 795029. [Google Scholar] [CrossRef] [PubMed]
  53. Darwich, N.F.; Phan, J.M.; Kim, B.; Suh, E.; Papatriantafyllou, J.D.; Changolkar, L.; Nguyen, A.T.; O’Rourke, C.M.; He, Z.; Porta, S.; et al. Autosomal dominant VCP hypomorph mutation impairs disaggregation of PHF-tau. Science 2020, 370, eaay8826. [Google Scholar] [CrossRef] [PubMed]
  54. Meinke, P.; Kerr, A.R.W.; Czapiewski, R.; de Las Heras, J.I.; Dixon, C.R.; Harris, E.; Kolbel, H.; Muntoni, F.; Schara, U.; Straub, V.; et al. A multistage sequencing strategy pinpoints novel candidate alleles for Emery-Dreifuss muscular dystrophy and supports gene misregulation as its pathomechanism. EBioMedicine 2019, 53, 102587. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Majounie, E.; Traynor, B.J.; Chio, A.; Restagno, G.; Mandrioli, J.; Benatar, M.; Taylor, J.P.; Singleton, A.B. Mutational analysis of the VCP gene in Parkinson’s disease. Neurobiol. Aging 2012, 33, 209.e1–209.e2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Rohrer, J.D.; Warren, J.D.; Reiman, D.; Uphill, J.; Beck, J.; Collinge, J.; Rossor, M.N.; Isaacs, A.M.; Mead, S. A novel exon 2 I27V VCP variant is associated with dissimilar clinical syndromes. J. Neurol. 2011, 258, 1494–1496. [Google Scholar] [CrossRef] [Green Version]
  57. Pensato, V.; Magri, S.; Bella, E.D.; Tannorella, P.; Bersano, E.; Soraru, G.; Gatti, M.; Ticozzi, N.; Taroni, F.; Lauria, G.; et al. Sorting Rare ALS Genetic Variants by Targeted Re-Sequencing Panel in Italian Patients: OPTN, VCP, and SQSTM1 Variants Account for 3% of Rare Genetic Forms. J. Clin. Med. 2020, 9, 412. [Google Scholar] [CrossRef] [Green Version]
  58. Mariani, L.L.; Tesson, C.; Charles, P.; Cazeneuve, C.; Hahn, V.; Youssov, K.; Freeman, L.; Grabli, D.; Roze, E.; Noel, S.; et al. Expanding the Spectrum of Genes Involved in Huntington Disease Using a Combined Clinical and Genetic Approach. JAMA Neurol. 2016, 73, 1105–1114. [Google Scholar] [CrossRef]
  59. Inoue, M.; Iida, A.; Hayashi, S.; Mori-Yoshimura, M.; Nagaoka, A.; Yoshimura, S.; Shiraishi, H.; Tsujino, A.; Takahashi, Y.; Nonaka, I.; et al. Two novel VCP missense variants identified in Japanese patients with multisystem proteinopathy. Hum. Genome Var. 2018, 5, 9. [Google Scholar] [CrossRef]
  60. Falcao de Campos, C.; de Carvalho, M. Distal myopathy and rapidly progressive dementia associated with a novel mutation in the VCP gene: Expanding inclusion body myopathy with early-onset Paget disease and frontotemporal dementia spectrum. J. Clin. Neurosci. 2019, 64, 8–10. [Google Scholar] [CrossRef]
  61. Deng, J.; Wu, W.; Xie, Z.; Gang, Q.; Yu, M.; Liu, J.; Wang, Q.; Lv, H.; Zhang, W.; Huang, Y.; et al. Novel and Recurrent Mutations in a Cohort of Chinese Patients With Young-Onset Amyotrophic Lateral Sclerosis. Front. Neurosci. 2019, 13, 1289. [Google Scholar] [CrossRef] [PubMed]
  62. Abrahao, A.; Abath Neto, O.; Kok, F.; Zanoteli, E.; Santos, B.; Pinto, W.B.; Barsottini, O.G.; Oliveira, A.S.; Pedroso, J.L. One family, one gene and three phenotypes: A novel VCP (valosin-containing protein) mutation associated with myopathy with rimmed vacuoles, amyotrophic lateral sclerosis and frontotemporal dementia. J. Neurol. Sci. 2016, 368, 352–358. [Google Scholar] [CrossRef] [PubMed]
  63. Hubbers, C.U.; Clemen, C.S.; Kesper, K.; Boddrich, A.; Hofmann, A.; Kamarainen, O.; Tolksdorf, K.; Stumpf, M.; Reichelt, J.; Roth, U.; et al. Pathological consequences of VCP mutations on human striated muscle. Brain 2007, 130, 381–393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Neveling, K.; Feenstra, I.; Gilissen, C.; Hoefsloot, L.H.; Kamsteeg, E.J.; Mensenkamp, A.R.; Rodenburg, R.J.; Yntema, H.G.; Spruijt, L.; Vermeer, S.; et al. A post-hoc comparison of the utility of sanger sequencing and exome sequencing for the diagnosis of heterogeneous diseases. Hum. Mutat. 2013, 34, 1721–1726. [Google Scholar] [CrossRef] [PubMed]
  65. Crisp, M.J.; Mawuenyega, K.G.; Patterson, B.W.; Reddy, N.C.; Chott, R.; Self, W.K.; Weihl, C.C.; Jockel-Balsarotti, J.; Varadhachary, A.S.; Bucelli, R.C.; et al. In vivo kinetic approach reveals slow SOD1 turnover in the CNS. J. Clin. Investig. 2015, 125, 2772–2780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Kaleem, M.; Zhao, A.; Hamshere, M.; Myers, A.J. Identification of a novel valosin-containing protein polymorphism in late-onset Alzheimer’s disease. Neurodegener. Dis. 2007, 4, 376–381. [Google Scholar] [CrossRef]
  67. Gu, J.M.; Ke, Y.H.; Yue, H.; Liu, Y.J.; Zhang, Z.; Zhang, H.; Hu, W.W.; Wang, C.; He, J.W.; Hu, Y.Q.; et al. A novel VCP mutation as the cause of atypical IBMPFD in a Chinese family. Bone 2013, 52, 9–16. [Google Scholar] [CrossRef]
  68. Ando, T.; Nakamura, R.; Kuru, S.; Yokoi, D.; Atsuta, N.; Koike, H.; Suzuki, M.; Hara, K.; Iguchi, Y.; Harada, Y.; et al. The wide-ranging clinical and genetic features in Japanese families with valosin-containing protein proteinopathy. Neurobiol. Aging 2020, 100, 120.e1–120.e6. [Google Scholar] [CrossRef]
  69. Saracino, D.; Clot, F.; Camuzat, A.; Anquetil, V.; Hannequin, D.; Guyant-Marechal, L.; Didic, M.; Guillot-Noel, L.; Rinaldi, D.; Latouche, M.; et al. Novel VCP mutations expand the mutational spectrum of frontotemporal dementia. Neurobiol. Aging 2018, 72, 187.e11–187.e14. [Google Scholar] [CrossRef]
  70. Koppers, M.; van Blitterswijk, M.M.; Vlam, L.; Rowicka, P.A.; van Vught, P.W.; Groen, E.J.; Spliet, W.G.; Engelen-Lee, J.; Schelhaas, H.J.; de Visser, M.; et al. VCP mutations in familial and sporadic amyotrophic lateral sclerosis. Neurobiol. Aging 2012, 33, 837.e7–837.e13. [Google Scholar] [CrossRef]
  71. Boland-Freitas, R.; Graham, J.; Davis, M.; Geevasinga, N.; Vucic, S.; Ng, K. Late onset distal myopathy of the upper limbs due to p.Ile151Val mutation in VCP. Muscle Nerve 2016, 54, 165–166. [Google Scholar] [CrossRef] [PubMed]
  72. De Ridder, W.; Azmi, A.; Clemen, C.S.; Eichinger, L.; Hofmann, A.; Schroder, R.; Johnson, K.; Topf, A.; Straub, V.; De Jonghe, P.; et al. Multisystem proteinopathy due to a homozygous p.Arg159His VCP mutation: A tale of the unexpected. Neurology 2019, 94, e785–e796. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Matsubara, S.; Shimizu, T.; Komori, T.; Mori-Yoshimura, M.; Minami, N.; Hayashi, Y.K. Nuclear inclusions mimicking poly(A)-binding protein nuclear 1 inclusions in a case of inclusion body myopathy associated with Paget disease of bone and frontotemporal dementia with a novel mutation in the valosin-containing protein gene. Neuromuscul. Disord. 2016, 26, 436–440. [Google Scholar] [CrossRef] [PubMed]
  74. Shi, Z.; Liu, S.; Xiang, L.; Wang, Y.; Liu, M.; Liu, S.; Han, T.; Zhou, Y.; Wang, J.; Cai, L.; et al. Frontotemporal dementia-related gene mutations in clinical dementia patients from a Chinese population. J. Hum. Genet. 2016, 61, 1003–1008. [Google Scholar] [CrossRef]
  75. Al-Tahan, S.; Al-Obeidi, E.; Yoshioka, H.; Lakatos, A.; Weiss, L.; Grafe, M.; Palmio, J.; Wicklund, M.; Harati, Y.; Omizo, M.; et al. Novel valosin-containing protein mutations associated with multisystem proteinopathy. Neuromuscul. Disord. 2018, 28, 491–501. [Google Scholar] [CrossRef] [Green Version]
  76. Lucente, G.; Almendrote, M.; Ramos-Fransi, A.; Martinez-Pineiro, A.; Camano, P.; Ballester-Lopez, A.; Lucia, A.; Carrato, C.; Nogales-Gadea, G.; Coll-Canti, J. A novel mutation in the valosin-containing-protein gene found in a Spanish family. J. Neurol. Sci. 2018, 391, 112–113. [Google Scholar] [CrossRef]
  77. Palmio, J.; Sandell, S.; Suominen, T.; Penttila, S.; Raheem, O.; Hackman, P.; Huovinen, S.; Haapasalo, H.; Udd, B. Distinct distal myopathy phenotype caused by VCP gene mutation in a Finnish family. Neuromuscul. Disord. 2011, 21, 551–555. [Google Scholar] [CrossRef]
  78. DeJesus-Hernandez, M.; Desaro, P.; Johnston, A.; Ross, O.A.; Wszolek, Z.K.; Ertekin-Taner, N.; Graff-Radford, N.R.; Rademakers, R.; Boylan, K. Novel p.Ile151Val mutation in VCP in a patient of African American descent with sporadic ALS. Neurology 2011, 77, 1102–1103. [Google Scholar] [CrossRef]
  79. Ramos, E.M.; Dokuru, D.R.; Van Berlo, V.; Wojta, K.; Wang, Q.; Huang, A.Y.; Deverasetty, S.; Qin, Y.; van Blitterswijk, M.; Jackson, J.; et al. Genetic screening of a large series of North American sporadic and familial frontotemporal dementia cases. Alzheimers Dement. 2020, 16, 118–130. [Google Scholar] [CrossRef]
  80. Gonzalez-Perez, P.; Cirulli, E.T.; Drory, V.E.; Dabby, R.; Nisipeanu, P.; Carasso, R.L.; Sadeh, M.; Fox, A.; Festoff, B.W.; Sapp, P.C.; et al. Novel mutation in VCP gene causes atypical amyotrophic lateral sclerosis. Neurology 2012, 79, 2201–2208. [Google Scholar] [CrossRef]
  81. Stojkovic, T.; Hammouda el, H.; Richard, P.; Lopez de Munain, A.; Ruiz-Martinez, J.; Camano, P.; Laforet, P.; Penisson-Besnier, I.; Ferrer, X.; Lacour, A.; et al. Clinical outcome in 19 French and Spanish patients with valosin-containing protein myopathy associated with Paget’s disease of bone and frontotemporal dementia. Neuromuscul. Disord. 2009, 19, 316–323. [Google Scholar] [CrossRef] [PubMed]
  82. Kumar, K.R.; Needham, M.; Mina, K.; Davis, M.; Brewer, J.; Staples, C.; Ng, K.; Sue, C.M.; Mastaglia, F.L. Two Australian families with inclusion-body myopathy, Paget’s disease of bone and frontotemporal dementia: Novel clinical and genetic findings. Neuromuscul. Disord. 2010, 20, 330–334. [Google Scholar] [CrossRef] [PubMed]
  83. Komatsu, J.; Iwasa, K.; Yanase, D.; Yamada, M. Inclusion body myopathy with Paget disease of the bone and frontotemporal dementia associated with a novel G156S mutation in the VCP gene. Muscle Nerve 2013. [Google Scholar] [CrossRef] [PubMed]
  84. Pellerin, D.; Ellezam, B.; Korathanakhun, P.; Renaud, M.; Dicaire, M.J.; Pilote, L.; Levy, J.P.; Karamchandani, J.; Ducharme, S.; Massie, R.; et al. Multi-system proteinopathy associated with a VCP G156S mutation in a French Canadian family. Can. J. Neurol. Sci. 2020, 47, 412–415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Segawa, M.; Hoshi, A.; Naruse, H.; Kuroda, M.; Bujo, H.; Ugawa, Y. A patient with familial amyotrophic lateral sclerosis associated with a new valosin-containing protein (VCP) gene mutation. Rinsho Shinkeigaku 2015, 55, 914–920. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Djamshidian, A.; Schaefer, J.; Haubenberger, D.; Stogmann, E.; Zimprich, F.; Auff, E.; Zimprich, A. A novel mutation in the VCP gene (G157R) in a German family with inclusion-body myopathy with Paget disease of bone and frontotemporal dementia. Muscle Nerve 2009, 39, 389–391. [Google Scholar] [CrossRef] [PubMed]
  87. Ayaki, T.; Ito, H.; Fukushima, H.; Inoue, T.; Kondo, T.; Ikemoto, A.; Asano, T.; Shodai, A.; Fujita, T.; Fukui, S.; et al. Immunoreactivity of valosin-containing protein in sporadic amyotrophic lateral sclerosis and in a case of its novel mutant. Acta Neuropathol. Commun. 2014, 2, 172. [Google Scholar] [CrossRef] [Green Version]
  88. Hang Wong, T.; Pottier, C.; Hondius, D.; Meeter, L.H.H.; van Rooij, J.G.J.; Melhem, S.; Brain Bank, T.N.; van Minkelen, R.; van Duijn, C.M.; Rozemuller, A.J.M.; et al. Three VCP Mutations in Patients with Frontotemporal Dementia. J. Alzheimers Dis. 2018, 65, 1139–1146. [Google Scholar] [CrossRef]
  89. Haubenberger, D.; Bittner, R.E.; Rauch-Shorny, S.; Zimprich, F.; Mannhalter, C.; Wagner, L.; Mineva, I.; Vass, K.; Auff, E.; Zimprich, A. Inclusion body myopathy and Paget disease is linked to a novel mutation in the VCP gene. Neurology 2005, 65, 1304–1305. [Google Scholar] [CrossRef]
  90. van der Zee, J.; Pirici, D.; Van Langenhove, T.; Engelborghs, S.; Vandenberghe, R.; Hoffmann, M.; Pusswald, G.; Van den Broeck, M.; Peeters, K.; Mattheijssens, M.; et al. Clinical heterogeneity in 3 unrelated families linked to VCP p.Arg159His. Neurology 2009, 73, 626–632. [Google Scholar] [CrossRef]
  91. Bersano, A.; Del Bo, R.; Lamperti, C.; Ghezzi, S.; Fagiolari, G.; Fortunato, F.; Ballabio, E.; Moggio, M.; Candelise, L.; Galimberti, D.; et al. Inclusion body myopathy and frontotemporal dementia caused by a novel VCP mutation. Neurobiol. Aging 2009, 30, 752–758. [Google Scholar] [CrossRef] [PubMed]
  92. Chan, N.; Le, C.; Shieh, P.; Mozaffar, T.; Khare, M.; Bronstein, J.; Kimonis, V. Valosin-containing protein mutation and Parkinson’s disease. Parkinsonism Relat. Disord. 2012, 18, 107–109. [Google Scholar] [CrossRef] [PubMed]
  93. de Bot, S.T.; Schelhaas, H.J.; Kamsteeg, E.J.; van de Warrenburg, B.P. Hereditary spastic paraplegia caused by a mutation in the VCP gene. Brain 2012, 135, e223. [Google Scholar] [CrossRef] [PubMed]
  94. Johnson, J.O.; Mandrioli, J.; Benatar, M.; Abramzon, Y.; Van Deerlin, V.M.; Trojanowski, J.Q.; Gibbs, J.R.; Brunetti, M.; Gronka, S.; Wuu, J.; et al. Exome sequencing reveals VCP mutations as a cause of familial ALS. Neuron 2010, 68, 857–864. [Google Scholar] [CrossRef] [Green Version]
  95. Kotan, D.; Ozozen Ayas, Z.; Tunca, C.; Gungen, B.D.; Akcimen, F.; Basak, A.N. Phenotypic and genotypic features of patients diagnosed with ALS in the city of Sakarya, Turkey. Acta Neurol. Belg. 2020, 120, 1411–1418. [Google Scholar] [CrossRef]
  96. Watts, G.D.; Thomasova, D.; Ramdeen, S.K.; Fulchiero, E.C.; Mehta, S.G.; Drachman, D.A.; Weihl, C.C.; Jamrozik, Z.; Kwiecinski, H.; Kaminska, A.; et al. Novel VCP mutations in inclusion body myopathy associated with Paget disease of bone and frontotemporal dementia. Clin. Genet. 2007, 72, 420–426. [Google Scholar] [CrossRef]
  97. Figueroa-Bonaparte, S.; Hudson, J.; Barresi, R.; Polvikoski, T.; Williams, T.; Topf, A.; Harris, E.; Hilton-Jones, D.; Petty, R.; Willis, T.A.; et al. Mutational spectrum and phenotypic variability of VCP-related neurological disease in the UK. J. Neurol. Neurosurg. Psychiatry 2015, 87, 680–681. [Google Scholar] [CrossRef]
  98. Peyer, A.K.; Kinter, J.; Hench, J.; Frank, S.; Fuhr, P.; Thomann, S.; Fischmann, A.; Kneifel, S.; Camano, P.; Lopez de Munain, A.; et al. Novel valosin containing protein mutation in a Swiss family with hereditary inclusion body myopathy and dementia. Neuromuscul. Disord. 2013, 23, 149–154. [Google Scholar] [CrossRef]
  99. Spina, S.; Van Laar, A.D.; Murrell, J.R.; Hamilton, R.L.; Kofler, J.K.; Epperson, F.; Farlow, M.R.; Lopez, O.L.; Quinlan, J.; DeKosky, S.T.; et al. Phenotypic variability in three families with valosin-containing protein mutation. Eur. J. Neurol. 2013, 20, 251–258. [Google Scholar] [CrossRef] [Green Version]
  100. Levesque, S.; Auray-Blais, C.; Gravel, E.; Boutin, M.; Dempsey-Nunez, L.; Jacques, P.E.; Chenier, S.; Larue, S.; Rioux, M.F.; Al-Hertani, W.; et al. Diagnosis of late-onset Pompe disease and other muscle disorders by next-generation sequencing. Orphanet J. Rare Dis. 2016, 11, 8. [Google Scholar] [CrossRef]
  101. Abramzon, Y.; Johnson, J.O.; Scholz, S.W.; Taylor, J.P.; Brunetti, M.; Calvo, A.; Mandrioli, J.; Benatar, M.; Mora, G.; Restagno, G.; et al. Valosin-containing protein (VCP) mutations in sporadic amyotrophic lateral sclerosis. Neurobiol. Aging 2012, 33, 2231.e1–2231.e6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Liewluck, T.; Milone, M.; Mauermann, M.L.; Castro-Couch, M.; Cerhan, J.H.; Murthy, N.S. A novel VCP mutation underlies scapuloperoneal muscular dystrophy and dropped head syndrome featuring lobulated fibers. Muscle Nerve 2014, 50, 295–299. [Google Scholar] [CrossRef] [PubMed]
  103. Shi, Z.; Hayashi, Y.K.; Mitsuhashi, S.; Goto, K.; Kaneda, D.; Choi, Y.C.; Toyoda, C.; Hieda, S.; Kamiyama, T.; Sato, H.; et al. Characterization of the Asian myopathy patients with VCP mutations. Eur. J. Neurol. 2012, 19, 501–509. [Google Scholar] [CrossRef] [PubMed]
  104. Hirano, M.; Nakamura, Y.; Saigoh, K.; Sakamoto, H.; Ueno, S.; Isono, C.; Mitsui, Y.; Kusunoki, S. VCP gene analyses in Japanese patients with sporadic amyotrophic lateral sclerosis identify a new mutation. Neurobiol. Aging 2015, 36, 1604.e1–1604.e6. [Google Scholar] [CrossRef]
  105. Papadopoulos, C.; Kirchner, P.; Bug, M.; Grum, D.; Koerver, L.; Schulze, N.; Poehler, R.; Dressler, A.; Fengler, S.; Arhzaouy, K.; et al. VCP/p97 cooperates with YOD1, UBXD1 and PLAA to drive clearance of ruptured lysosomes by autophagy. EMBO J. 2017, 36, 135–150. [Google Scholar] [CrossRef] [Green Version]
  106. Kenna, K.P.; McLaughlin, R.L.; Byrne, S.; Elamin, M.; Heverin, M.; Kenny, E.M.; Cormican, P.; Morris, D.W.; Donaghy, C.G.; Bradley, D.G.; et al. Delineating the genetic heterogeneity of ALS using targeted high-throughput sequencing. J. Med. Genet. 2013, 50, 776–783. [Google Scholar] [CrossRef] [Green Version]
  107. Iossifov, I.; Ronemus, M.; Levy, D.; Wang, Z.; Hakker, I.; Rosenbaum, J.; Yamrom, B.; Lee, Y.H.; Narzisi, G.; Leotta, A.; et al. De novo gene disruptions in children on the autistic spectrum. Neuron 2012, 74, 285–299. [Google Scholar] [CrossRef] [Green Version]
  108. 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]
  109. Schrodinger, L.; DeLano, W. PyMOL. Available online: http://www.pymol.org/pymol (accessed on 24 April 2022).
  110. Pan, M.; Yu, Y.; Ai, H.; Zheng, Q.; Xie, Y.; Liu, L.; Zhao, M. Mechanistic insight into substrate processing and allosteric inhibition of human p97. Nat. Struct. Mol. Biol. 2021, 28, 614–625. [Google Scholar] [CrossRef]
  111. Giordano, L.; Deceglie, S.; d’Adamo, P.; Valentino, M.L.; La Morgia, C.; Fracasso, F.; Roberti, M.; Cappellari, M.; Petrosillo, G.; Ciaravolo, S.; et al. Cigarette toxicity triggers Leber’s hereditary optic neuropathy by affecting mtDNA copy number, oxidative phosphorylation and ROS detoxification pathways. Cell Death Dis. 2015, 6, e2021. [Google Scholar] [CrossRef] [Green Version]
  112. Pfeffer, G.; Cote, H.C.; Montaner, J.S.; Li, C.C.; Jitratkosol, M.; Mezei, M.M. Ophthalmoplegia and ptosis: Mitochondrial toxicity in patients receiving HIV therapy. Neurology 2009, 73, 71–72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Pfeffer, G.; Sirrs, S.; Wade, N.K.; Mezei, M.M. Multisystem disorder in late-onset chronic progressive external ophthalmoplegia. Can. J. Neurol. Sci. 2011, 38, 119–123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Blacher, E.; Bashiardes, S.; Shapiro, H.; Rothschild, D.; Mor, U.; Dori-Bachash, M.; Kleimeyer, C.; Moresi, C.; Harnik, Y.; Zur, M.; et al. Potential roles of gut microbiome and metabolites in modulating ALS in mice. Nature 2019, 572, 474–480. [Google Scholar] [CrossRef] [PubMed]
  115. Mehta, S.G.; Watts, G.D.; Adamson, J.L.; Hutton, M.; Umberger, G.; Xiong, S.; Ramdeen, S.; Lovell, M.A.; Kimonis, V.E.; Smith, C.D. APOE is a potential modifier gene in an autosomal dominant form of frontotemporal dementia (IBMPFD). Genet. Med. 2007, 9, 9–13. [Google Scholar] [CrossRef] [Green Version]
  116. Kirby, A.E.; Kimonis, V.; Kompoliti, K. Ataxia and Parkinsonism in a Woman With a VCP Variant and Long-Normal Repeats in the SCA2 Allele. Neurol. Genet. 2021, 7, e595. [Google Scholar] [CrossRef]
  117. Chalmers, M.R.; Kim, J.; Kim, N.C. Eip74EF is a dominant modifier for ALS-FTD linked mutant VCP phenotypes in Drosophila, but not miR-34. Biorxiv 2020. [Google Scholar] [CrossRef]
  118. Ikenaga, C.; Findlay, A.R.; Goyal, N.A.; Robinson, S.; Cauchi, J.; Hussain, Y.; Wang, L.H.; Kershen, J.C.; Beson, B.A.; Wallendorf, M.; et al. Clinical utility of anti-cytosolic 5’-nucleotidase 1A antibody in idiopathic inflammatory myopathies. Ann. Clin. Transl. Neurol. 2021, 8, 571–578. [Google Scholar] [CrossRef]
  119. Posey, J.E.; Harel, T.; Liu, P.; Rosenfeld, J.A.; James, R.A.; Coban Akdemir, Z.H.; Walkiewicz, M.; Bi, W.; Xiao, R.; Ding, Y.; et al. Resolution of Disease Phenotypes Resulting from Multilocus Genomic Variation. N. Engl. J. Med. 2017, 376, 21–31. [Google Scholar] [CrossRef]
  120. Estiar, M.A.; Yu, E.; Haj Salem, I.; Ross, J.P.; Mufti, K.; Akcimen, F.; Leveille, E.; Spiegelman, D.; Ruskey, J.A.; Asayesh, F.; et al. Evidence for Non-Mendelian Inheritance in Spastic Paraplegia 7. Mov. Disord. 2021, 36, 1664–1675. [Google Scholar] [CrossRef]
  121. Weihl, C.C.; Pestronk, A.; Kimonis, V.E. Valosin-containing protein disease: Inclusion body myopathy with Paget’s disease of the bone and fronto-temporal dementia. Neuromuscul. Disord. 2009, 19, 308–315. [Google Scholar] [CrossRef] [Green Version]
  122. Bugiardini, E.; Morrow, J.M.; Shah, S.; Wood, C.L.; Lynch, D.S.; Pitmann, A.M.; Reilly, M.M.; Houlden, H.; Matthews, E.; Parton, M.; et al. The Diagnostic Value of MRI Pattern Recognition in Distal Myopathies. Front. Neurol. 2018, 9, 456. [Google Scholar] [CrossRef] [PubMed]
  123. Diaz-Manera, J.; Llauger, J.; Gallardo, E.; Illa, I. Muscle MRI in muscular dystrophies. Acta Myol. 2015, 34, 95–108. [Google Scholar] [PubMed]
  124. Papadimas, G.K.; Paraskevas, G.P.; Zambelis, T.; Karagiaouris, C.; Bourbouli, M.; Bougea, A.; Walter, M.C.; Schumacher, N.U.; Krause, S.; Kapaki, E. The multifaceted clinical presentation of VCP-proteinopathy in a Greek family. Acta Myol. 2017, 36, 203–206. [Google Scholar] [PubMed]
  125. Paul Tuck, S.; Layfield, R.; Walker, J.; Mekkayil, B.; Francis, R. Adult Paget’s disease of bone: A review. Rheumatology 2017, 56, 2050–2059. [Google Scholar] [CrossRef] [Green Version]
  126. Richards, S.; Aziz, N.; Bale, S.; Bick, D.; Das, S.; Gastier-Foster, J.; Grody, W.W.; Hegde, M.; Lyon, E.; Spector, E.; et al. Standards and guidelines for the interpretation of sequence variants: A joint consensus recommendation of the American College of Medical Genetics and Genomics and the Association for Molecular Pathology. Genet. Med. 2015, 17, 405–424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Grimes, D.A.; Barclay, C.L.; Duff, J.; Furukawa, Y.; Lang, A.E. Phenocopies in a large GCH1 mutation positive family with dopa responsive dystonia: Confusing the picture? J. Neurol. Neurosurg. Psychiatry 2002, 72, 801–804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Garrison, N.A.; Brothers, K.B.; Goldenberg, A.J.; Lynch, J.A. Genomic Contextualism: Shifting the Rhetoric of Genetic Exceptionalism. Am. J. Bioeth. 2019, 19, 51–63. [Google Scholar] [CrossRef] [Green Version]
  129. Laberge, A.M.; Richer, J.; Ravitsky, V. Toward Broader Genetic Contextualism: Genetic Testing Enters the Age of Evidence-Based Medicine. Am. J. Bioeth. 2019, 19, 77–79. [Google Scholar] [CrossRef]
  130. Slade, I.; Riddell, D.; Turnbull, C.; Hanson, H.; Rahman, N.; programme, M.C.G. Development of cancer genetic services in the UK: A national consultation. Genome Med. 2015, 7, 18. [Google Scholar] [CrossRef] [Green Version]
  131. Cohen, S.A.; Bradbury, A.; Henderson, V.; Hoskins, K.; Bednar, E.; Arun, B.K. Genetic Counseling and Testing in a Community Setting: Quality, Access, and Efficiency. Am. Soc. Clin. Oncol. Educ. Book 2019, 39, e33–e44. [Google Scholar] [CrossRef]
  132. Yuen, J.; Cousens, N.; Barlow-Stewart, K.; O’Shea, R.; Andrews, L. Online BRCA1/2 screening in the Australian Jewish community: A qualitative study. J. Community Genet. 2020, 11, 291–302. [Google Scholar] [CrossRef] [PubMed]
  133. Scuffham, T.M.; MacMillan, J.C. Huntington disease: Who seeks presymptomatic genetic testing, why and what are the outcomes? J. Genet. Couns. 2014, 23, 754–761. [Google Scholar] [CrossRef] [PubMed]
  134. Godino, L.; Turchetti, D.; Jackson, L.; Hennessy, C.; Skirton, H. Impact of presymptomatic genetic testing on young adults: A systematic review. Eur. J. Hum. Genet. 2016, 24, 496–503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Tilleras, K.H.; Kjoelaas, S.H.; Dramstad, E.; Feragen, K.B.; von der Lippe, C. Psychological reactions to predictive genetic testing for Huntington’s disease: A qualitative study. J. Genet. Couns. 2020, 29, 1093–1105. [Google Scholar] [CrossRef] [Green Version]
  136. Surampalli, A.; Khare, M.; Kubrussi, G.; Wencel, M.; Tanaja, J.; Donkervoort, S.; Osann, K.; Simon, M.; Wallace, D.; Smith, C.; et al. Psychological Impact of Predictive Genetic Testing in VCP Inclusion Body Myopathy, Paget Disease of Bone and Frontotemporal Dementia. J. Genet. Couns. 2015, 24, 842–850. [Google Scholar] [CrossRef] [Green Version]
  137. Clement, S.; Gargiulo, M.; Feingold, J.; Durr, A. Guidelines for presymptomatic testing for Huntington’s disease: Past, present and future in France. Rev. Neurol. 2015, 171, 572–580. [Google Scholar] [CrossRef]
  138. Dalpe, G.; Pinkesz, M.; Oliviero, E.; Tolymbek, M.; Joly, Y. Genetic discrimination views in online discussion forums: Perspectives from Canadian forumites. J. Genet. Couns. 2021, 30, 1613–1628. [Google Scholar] [CrossRef]
  139. Burnett-Hartman, A.N.; Blum-Barnett, E.; Carroll, N.M.; Madrid, S.D.; Jonas, C.; Janes, K.; Alvarado, M.; Bedoy, R.; Paolino, V.; Aziz, N.; et al. Return of Research-Related Genetic Test Results and Genetic Discrimination Concerns: Facilitators and Barriers of Genetic Research Participation in Diverse Groups. Public Health Genomics. 2020, 23, 59–68. [Google Scholar] [CrossRef]
  140. Joly, Y.; Dalpe, G.; Dupras, C.; Beviere-Boyer, B.; de Paor, A.; Dove, E.S.; Granados Moreno, P.; Ho, C.W.L.; Ho, C.H.; K, O.C.; et al. Establishing the International Genetic Discrimination Observatory. Nat. Genet. 2020, 52, 466–468. [Google Scholar] [CrossRef]
  141. de Paor, A. Genetic Discrimination: A Case for a European Legislative Response? Eur. J. Health Law 2017, 24, 135–159. [Google Scholar] [CrossRef]
  142. Clifton, J.M.; VanBeuge, S.S.; Mladenka, C.; Wosnik, K.K. The Genetic Information Nondiscrimination Act 2008: What clinicians should understand. J. Am. Acad. Nurse Pract. 2010, 22, 246–249. [Google Scholar] [CrossRef] [PubMed]
Table 1. Neurological localizations and clinical phenotypes associated with VCP-MSP.
Table 1. Neurological localizations and clinical phenotypes associated with VCP-MSP.
LocalizationPhenotype or DiseaseFrequency in VCP-MSP
Central nervous systemFrontotemporal dementia30%
Alzheimer disease2%
Parkinson disease4%
Spastic paraplegiaCase reports
Peripheral nervous systemInclusion body myopathy90%
Amyotrophic lateral sclerosis10%
Charcot-Marie-Tooth diseaseCase reports
Sensory polyneuropathyCase reports
Non-neurologicalPaget disease of bone40%
CardiomyopathyUncertain
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pfeffer, G.; Lee, G.; Pontifex, C.S.; Fanganiello, R.D.; Peck, A.; Weihl, C.C.; Kimonis, V. Multisystem Proteinopathy Due to VCP Mutations: A Review of Clinical Heterogeneity and Genetic Diagnosis. Genes 2022, 13, 963. https://doi.org/10.3390/genes13060963

AMA Style

Pfeffer G, Lee G, Pontifex CS, Fanganiello RD, Peck A, Weihl CC, Kimonis V. Multisystem Proteinopathy Due to VCP Mutations: A Review of Clinical Heterogeneity and Genetic Diagnosis. Genes. 2022; 13(6):963. https://doi.org/10.3390/genes13060963

Chicago/Turabian Style

Pfeffer, Gerald, Grace Lee, Carly S. Pontifex, Roberto D. Fanganiello, Allison Peck, Conrad C. Weihl, and Virginia Kimonis. 2022. "Multisystem Proteinopathy Due to VCP Mutations: A Review of Clinical Heterogeneity and Genetic Diagnosis" Genes 13, no. 6: 963. https://doi.org/10.3390/genes13060963

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