Entry - #182601 - SPASTIC PARAPLEGIA 4, AUTOSOMAL DOMINANT; SPG4 - OMIM
# 182601

SPASTIC PARAPLEGIA 4, AUTOSOMAL DOMINANT; SPG4


Alternative titles; symbols

FAMILIAL SPASTIC PARAPLEGIA, AUTOSOMAL DOMINANT, 2; FSP2


Phenotype-Gene Relationships

Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
Gene/Locus Gene/Locus
MIM number
2p22.3 Spastic paraplegia 4, autosomal dominant 182601 AD 3 SPAST 604277
Clinical Synopsis
 
Phenotypic Series
 

INHERITANCE
- Autosomal dominant
HEAD & NECK
Eyes
- Nystagmus (rare)
GENITOURINARY
Bladder
- Urinary urgency
- Urinary incontinence
- Sphincter disturbances
SKELETAL
Spine
- Lower back pain
NEUROLOGIC
Central Nervous System
- Lower limb spasticity
- Lower limb weakness
- Spastic gait
- Hyperreflexia
- Extensor plantar responses
- Pyramidal signs
- Degeneration of the lateral corticospinal tracts
- Cognitive decline
- Memory impairment
- Deficits in language expression
- Deficits in abstraction
- Mental retardation (rare)
- Dementia (rare)
- Arachnoid cysts of the cerebellopontine angle (reported in 1 family)
Peripheral Nervous System
- Decreased vibratory sense in the lower limbs
Behavioral Psychiatric Manifestations
- Agitation
- Aggression
- Apathy
- Depression
- Disinhibition
MISCELLANEOUS
- Variable age of onset (infancy to 63 years)
- Insidious onset
- Progressive disorder
- Highly variable severity
- Genetic anticipation
- Most common form of autosomal dominant hereditary spastic paraplegia (accounts for 40% of SPG cases)
- Genetic heterogeneity, see SPG3A (182600)
MOLECULAR BASIS
- Caused by mutation in the spastin gene (SPG4, 604277.0001)
Spastic paraplegia - PS303350 - 83 Entries
Location Phenotype Inheritance Phenotype
mapping key
Phenotype
MIM number
Gene/Locus Gene/Locus
MIM number
1p36.13 Spastic paraplegia 78, autosomal recessive AR 3 617225 ATP13A2 610513
1p34.1 Spastic paraplegia 83, autosomal recessive AR 3 619027 HPDL 618994
1p31.1-p21.1 Spastic paraplegia 29, autosomal dominant AD 2 609727 SPG29 609727
1p13.3 ?Spastic paraplegia 63, autosomal recessive AR 3 615686 AMPD2 102771
1p13.2 Spastic paraplegia 47, autosomal recessive AR 3 614066 AP4B1 607245
1q32.1 Spastic paraplegia 23, autosomal recessive AR 3 270750 DSTYK 612666
1q42.13 ?Spastic paraplegia 44, autosomal recessive AR 3 613206 GJC2 608803
1q42.13 ?Spastic paraplegia 74, autosomal recessive AR 3 616451 IBA57 615316
2p23.3 Spastic paraplegia 81, autosomal recessive AR 3 618768 SELENOI 607915
2p22.3 Spastic paraplegia 4, autosomal dominant AD 3 182601 SPAST 604277
2p11.2 Spastic paraplegia 31, autosomal dominant AD 3 610250 REEP1 609139
2q33.1 Spastic paraplegia 13, autosomal dominant AD 3 605280 HSPD1 118190
2q37.3 Spastic paraplegia 30, autosomal recessive AD, AR 3 610357 KIF1A 601255
2q37.3 Spastic paraplegia 30, autosomal dominant AD, AR 3 610357 KIF1A 601255
3q12.2 ?Spastic paraplegia 57, autosomal recessive AR 3 615658 TFG 602498
3q25.31 Spastic paraplegia 42, autosomal dominant AD 3 612539 SLC33A1 603690
3q27-q28 Spastic paraplegia 14, autosomal recessive AR 2 605229 SPG14 605229
4p16-p15 Spastic paraplegia 38, autosomal dominant AD 2 612335 SPG38 612335
4p13 Spastic paraplegia 79B, autosomal recessive AR 3 615491 UCHL1 191342
4p13 Spastic paraplegia 79A, autosomal dominant AD 3 620221 UCHL1 191342
4q25 Spastic paraplegia 56, autosomal recessive AR 3 615030 CYP2U1 610670
5q31.2 ?Spastic paraplegia 72B, autosomal recessive AR 3 620606 REEP2 609347
5q31.2 Spastic paraplegia 72A, autosomal dominant AD 3 615625 REEP2 609347
6p25.1 Spastic paraplegia 77, autosomal recessive AR 3 617046 FARS2 611592
6p21.33 Spastic paraplegia 86, autosomal recessive AR 3 619735 ABHD16A 142620
6q23-q24.1 Spastic paraplegia 25, autosomal recessive AR 2 608220 SPG25 608220
7p22.1 Spastic paraplegia 48, autosomal recessive AR 3 613647 AP5Z1 613653
7q22.1 Spastic paraplegia 50, autosomal recessive AR 3 612936 AP4M1 602296
8p22 Spastic paraplegia 53, autosomal recessive AR 3 614898 VPS37A 609927
8p21.1-q13.3 Spastic paraplegia 37, autosomal dominant AD 2 611945 SPG37 611945
8p11.23 Spastic paraplegia 18B, autosomal recessive AR 3 611225 ERLIN2 611605
8p11.23 Spastic paraplegia 18A, autosomal dominant AD 3 620512 ERLIN2 611605
8p11.23 Spastic paraplegia 54, autosomal recessive AR 3 615033 DDHD2 615003
8p11.21 Spastic paraplegia 85, autosomal recessive AR 3 619686 RNF170 614649
8q12.3 Spastic paraplegia 5A, autosomal recessive AR 3 270800 CYP7B1 603711
8q24.13 Spastic paraplegia 8, autosomal dominant AD 3 603563 WASHC5 610657
9p13.3 Spastic paraplegia 46, autosomal recessive AR 3 614409 GBA2 609471
9q Spastic paraplegia 19, autosomal dominant AD 2 607152 SPG19 607152
10q22.1-q24.1 Spastic paraplegia 27, autosomal recessive AR 2 609041 SPG27 609041
10q24.1 Spastic paraplegia 9B, autosomal recessive AR 3 616586 ALDH18A1 138250
10q24.1 Spastic paraplegia 9A, autosomal dominant AD 3 601162 ALDH18A1 138250
10q24.1 Spastic paraplegia 64, autosomal recessive AR 3 615683 ENTPD1 601752
10q24.2 Spastic paraplegia 33, autosomal dominant AD 3 610244 ZFYVE27 610243
10q24.31 Spastic paraplegia 62, autosomal recessive AR 3 615681 ERLIN1 611604
10q24.32-q24.33 Spastic paraplegia 45, autosomal recessive AR 3 613162 NT5C2 600417
11p14.1-p11.2 ?Spastic paraplegia 41, autosomal dominant AD 2 613364 SPG41 613364
11q12.3 Silver spastic paraplegia syndrome AD 3 270685 BSCL2 606158
11q13.1 Spastic paraplegia 76, autosomal recessive AR 3 616907 CAPN1 114220
12q13.3 Spastic paraplegia 70, autosomal recessive AR 3 620323 MARS1 156560
12q13.3 Spastic paraplegia 10, autosomal dominant AD 3 604187 KIF5A 602821
12q13.3 Spastic paraplegia 26, autosomal recessive AR 3 609195 B4GALNT1 601873
12q23-q24 Spastic paraplegia 36, autosomal dominant AD 2 613096 SPG36 613096
12q24.31 Spastic paraplegia 55, autosomal recessive AR 3 615035 MTRFR 613541
13q13.3 Troyer syndrome AR 3 275900 SPART 607111
13q14 Spastic paraplegia 24, autosomal recessive AR 2 607584 SPG24 607584
13q14.2 Spastic paraplegia 88, autosomal dominant AD 3 620106 KPNA3 601892
14q12-q21 Spastic paraplegia 32, autosomal recessive AR 2 611252 SPG32 611252
14q12 Spastic paraplegia 52, autosomal recessive AR 3 614067 AP4S1 607243
14q13.1 Spastic paraplegia 90A, autosomal dominant AD 3 620416 SPTSSA 613540
14q13.1 ?Spastic paraplegia 90B, autosomal recessive AD 3 620417 SPTSSA 613540
14q22.1 Spastic paraplegia 3A, autosomal dominant AD 3 182600 ATL1 606439
14q22.1 Spastic paraplegia 28, autosomal recessive AR 3 609340 DDHD1 614603
14q24.1 Spastic paraplegia 15, autosomal recessive AR 3 270700 ZFYVE26 612012
14q24.3 Spastic paraplegia 87, autosomal recessive AR 3 619966 TMEM63C 619953
15q11.2 Spastic paraplegia 6, autosomal dominant AD 3 600363 NIPA1 608145
15q21.1 Spastic paraplegia 11, autosomal recessive AR 3 604360 SPG11 610844
15q21.2 Spastic paraplegia 51, autosomal recessive AR 3 613744 AP4E1 607244
15q22.31 Mast syndrome AR 3 248900 ACP33 608181
16p12.3 Spastic paraplegia 61, autosomal recessive AR 3 615685 ARL6IP1 607669
16q13 Spastic paraplegia 89, autosomal recessive AR 3 620379 AMFR 603243
16q23.1 Spastic paraplegia 35, autosomal recessive AR 3 612319 FA2H 611026
16q24.3 Spastic paraplegia 7, autosomal recessive AD, AR 3 607259 PGN 602783
17q25.3 Spastic paraplegia 82, autosomal recessive AR 3 618770 PCYT2 602679
19p13.2 Spastic paraplegia 39, autosomal recessive AR 3 612020 PNPLA6 603197
19q12 ?Spastic paraplegia 43, autosomal recessive AR 3 615043 C19orf12 614297
19q13.12 Spastic paraplegia 75, autosomal recessive AR 3 616680 MAG 159460
19q13.32 Spastic paraplegia 12, autosomal dominant AD 3 604805 RTN2 603183
19q13.33 ?Spastic paraplegia 73, autosomal dominant AD 3 616282 CPT1C 608846
22q11.21 Spastic paraplegia 84, autosomal recessive AR 3 619621 PI4KA 600286
Xq11.2 Spastic paraplegia 16, X-linked, complicated XLR 2 300266 SPG16 300266
Xq22.2 Spastic paraplegia 2, X-linked XLR 3 312920 PLP1 300401
Xq24-q25 Spastic paraplegia 34, X-linked XLR 2 300750 SPG34 300750
Xq28 MASA syndrome XLR 3 303350 L1CAM 308840

TEXT

A number sign (#) is used with this entry because autosomal dominant spastic paraplegia-4 (SPG4) is caused by heterozygous mutation in the SPAST gene (604277) on chromosome 2p22.


Description

The hereditary spastic paraplegias (SPG, HSP) are a group of clinically and genetically diverse inherited disorders characterized predominantly by progressive lower extremity spasticity and weakness. SPG is classified by mode of inheritance (autosomal dominant, autosomal recessive, and X-linked) and whether the primary symptoms occur in isolation ('uncomplicated') or with other neurologic abnormalities ('complicated').

Pure SPG4 is the most common form of autosomal dominant hereditary SPG, comprising up to 45% of cases (Svenson et al., 2001; Crippa et al., 2006).

For a general phenotypic description and a discussion of genetic heterogeneity of autosomal dominant spastic paraplegia, see SPG3A (182600).


Clinical Features

In 5 of 7 French families and in 1 large Dutch pedigree with a form of autosomal dominant familial spastic paraplegia, Hazan et al. (1994) found linkage to a locus, which they termed FSP2 (also known as SPG4), on chromosome 2p. This finding distinguished the disease from autosomal dominant spastic paraplegia-3 (182600), which had been mapped to chromosome 14. Age of onset in the 6 families showing linkage to 2p varied widely within families and the mean age at onset ranged from 20 to 39 years. Thus, age of onset may be a poor criterion for classifying autosomal dominant spastic paraplegia. Anticipation in the age of onset was observed in 2 of the kindreds.

Durr et al. (1996) reported 12 families with autosomal dominant spastic paraplegia linked to the SPG4 locus on chromosome 2. Age of onset ranged from infancy to 63 years. The clinical expression of the disorder within a family included asymptomatic patients who were unaware of their condition, mildly affected individuals who had spastic gait but were able to walk independently, and severely affected patients who were wheelchair bound. Durr et al. (1996) commented on the extensive intra- and interfamilial clinical variation.

Nielsen et al. (1998) evaluated 5 families with 2p-linked pure spastic paraplegia. In 2 families, nonprogressive 'congenital' spastic paraplegia was seen in some affected members, whereas adult-onset progressive spastic paraplegia was present in others. Low backache was reported as a late symptom by 47% of the 63 at-risk members in the 5 families. Brain and total spinal cord MRI disclosed no significant abnormalities. Nielsen et al. (1998) concluded that SPG4 is a phenotypically heterogeneous disorder, characterized by both interfamilial and intrafamilial variation.

Nance et al. (1998) found striking variation in clinical features in 4 families with spastic paraplegia with linkage to chromosome 2 markers. Only mild neurologic signs were observed in some subjects. The clinical features of 1 family had previously been described by Boustany et al. (1987). Onset was generally in the third to fifth decades with an average onset age of 35 years (range, 5 to 61 years). All clearly affected patients had scissoring gait, and in all who were examined at least 2 of the following were found: extensor plantar responses, increased knee and ankle reflexes, increased tone, muscle spasms, or leg cramps. Urinary urgency or other symptoms compatible with a neurogenic bladder, leg weakness, and decreased vibration sense were present in some, but not all, patients.

Byrne et al. (1997, 1998) presented a family with autosomal dominant hereditary spastic paraplegia and a specific form of cognitive impairment who showed linkage to the SPG4 locus on chromosome 2. The pattern of cognitive impairment in this family was characterized primarily by deficits in visual-spatial functions. Dysfunction manifested itself by difficulty in carrying out new tasks, forgetfulness, poor spatial perception, and impaired visual-motor coordination. By haplotype analysis the presence of the mutant gene was identified in an individual who, at the age of 57, had the same pattern of cognitive impairment but no spastic paraplegia. Furthermore, 6 individuals who presented with the disease haplotype had normal neurologic and neuropsychologic examinations. All 6 were below the maximal age of onset in the family, namely, 60 years. In this Irish family the cognitive impairment was considered to be a manifestation of the SPG4 gene mutation.

Reid et al. (1999) investigated 35 individuals from 4 families of Welsh origin, 22 of whom had 'pure' hereditary spastic paraplegia, for the presence of subclinical cognitive impairment. They found significant reductions in scores on the Mini-Mental State Examination (MMSE) among affected individuals compared to controls. To assess whether the lower MMSE scores were restricted to subjects older than 50 years, scores for affected subjects 50 years of age or younger were compared to those of controls. A significant difference in score remained. One of the families was linked to the chromosome 2 locus, while 2 others showed linkage to none of the loci known at that time. There was no significant difference between the results of these 2 groups.

McMonagle et al. (2000) compared the phenotypic expressions of autosomal dominant hereditary spastic paresis in several families with a mutation in the SPG4 gene and several families without a mutation in SPG4. In the mutation-positive group, age of onset was later, disability score was greater, progression of disease was faster, wheelchair use was greater (40.9% vs 4.8% in the mutation-excluded group), there was greater abnormal vibration sensation in the lower limbs (68.2% vs 19%), and fewer individuals were asymptomatic (18.2% vs 42.9%). Dementia was more prevalent in the mutation-positive group. McMonagle et al. (2000) emphasized the finding of cognitive impairment as a feature of SPG4 mutations.

White et al. (2000) reported a patient with familial SPG4 who had clinical dementia. Postmortem neuropathologic examination showed neuronal loss and tau- (MAPT; 157140) immunoreactive neurofibrillary tangles in the hippocampus and tau-immunoreactive balloon cells in the limbic area and neocortex. Lewy bodies were present in the substantia nigra. White et al. (2000) suggested that these findings confirmed an association of dementia with SPG4.

McMonagle et al. (2004) used several measures of cognitive function to assess 11 patients from 3 families in whom SPG4 was confirmed by genetic analysis or linkage. SPG4 patients scored significantly lower on the Cambridge cognitive examination (CAMCOG) (mean score of 73.5 compared to 91.7 in controls). After approximately 3 years, the patients' mean score fell to 64.4, whereas the mean control score declined slightly to 90.8. Deficits in the SPG4 patients were noted in attention, language expression, memory, and abstraction. Behavior assessment found that SPG4 patients exhibited agitation, aggression, apathy, irritability, depression, and disinhibition. Accounting for age, McMonagle et al. (2004) concluded that subtle changes in cognitive function in patients with SPG4 may begin after age 40 years, with more severe decline after age 60.

Orlacchio et al. (2004) reported 32 patients from 9 families from southern Scotland with SPG4. Age at onset varied from 11 to 53 years. In addition to classic features of hereditary spastic paraplegia, 2 of the 32 patients had mental retardation and 2 other patients had a thin corpus callosum and cerebellar atrophy. All affected members had the same mutation in the SPG4 gene (604277.0014), and haplotype analysis suggested a founder effect.

Orlacchio et al. (2004) reported a large Italian family in which all 16 members who had SPG4 also had congenital arachnoid cysts at the cerebellopontine angle ranging in size from 21 to 31 mm. Six patients also had mental retardation. Genetic analysis confirmed a mutation in the SPG4 gene.

McDermott et al. (2006) reported a patient with SPG4 who developed walking difficulties in his late teens with deteriorating gait in his 20s; he was wheelchair-dependent at age 35. He later developed stiffness in the upper limbs, bladder dysfunction, dysarthria, and swallowing difficulties. In his 40s, he developed respiratory insufficiency and distal muscle wasting in the lower limbs. Molecular analysis identified a mutation in the SPG4 gene (S445R; 604277.0021). The findings of bulbar and respiratory involvement, as well as lower motor neuron degeneration, broadened the phenotype associated with mutations in the SPG4 gene.

Orlacchio et al. (2008) reported a large 4-generation Italian family with SPG4 confirmed by genetic analysis. The mean ages at onset were 17.5 and 18.8 years for symptoms of the lower and upper limbs, respectively. There was a general impression of genetic anticipation spanning the 4 generations. All affected individuals had spasticity of the lower limbs and pyramidal tract signs such as hyperreflexia, extensor plantar responses, or both, and pes cavus. All patients also had weak intrinsic hand muscles, with severe amyotrophy most relevant in the thenar eminence. Peroneal muscle wasting was reported in five patients, and many used a cane. Other associated features included impaired vibration sensation and cognitive dysfunctions. All patients except 1 had temporal lobe epilepsy with partial complex seizures associated with hippocampal sclerosis.

Murphy et al. (2009) reported a family in which 12 members had SPG4 due to a deletion of exon 17 in the SPG4 gene (Beetz et al., 2007). Cognitive assessment performed over a 7-year period found that all 4 patients who were older than 60 years developed mild to moderate cognitive decline. Two younger patients aged 48 and 40, respectively, had mild cognitive impairment. Genetic analysis of this family was unusual because 4 patients with the SPG4 deletion also carried a microdeletion in the NIPA1 gene (608145), which causes SPG6 (600363); only 2 of these 4 had cognitive impairment. Five patients with only the SPG4 deletion had cognitive impairment, including 2 who did not have clinical signs of SPG. Another family member with only the NIPA1 microdeletion lacked clinical signs of SPG or cognitive impairment at age 57. Murphy et al. (2009) concluded that SPG4 is associated with cognitive decline, and that the SPG6 microdeletion does not have a clinical phenotype in this family. Postmortem examination of the proband, who had both deletions as well as SPG and cognitive impairment, showed a markedly atrophic spinal cord with degeneration of the corticospinal tracts, and superficial spongiosis and widespread ubiquitin-positive inclusions in the neocortex and white matter.


Inheritance

The transmission pattern of SPG4 in the families reported by Hazan et al. (1999) was consistent with autosomal dominant inheritance.


Mapping

In 5 French families and 1 large Dutch pedigree with autosomal dominant spastic paraplegia, Hazan et al. (1994) found linkage markers in the 2p24-p21 region. An analysis of recombination events and multipoint linkage placed this form of the disease within a 4-cM interval flanked by loci D2S400 and D2S367. In 4 Caucasian North American families and in 1 family from Tunisia, Hentati et al. (1994) found linkage of late-onset SPG to DNA markers on chromosome 2p in 4 of the families. Pathologic findings in a member of one of the chromosome 2-linked families had previously been reported by Sack et al. (1978).

Scott et al. (1997) examined 11 Caucasian pedigrees with autosomal dominant 'uncomplicated' familial spastic paraplegia for linkage to the previously identified loci on 2p, 14q (SPG3A), and 15q (SPG6; 600363). Chromosome 15q was excluded for all families. Five families showed evidence for linkage to 2p, 1 family to 14q, and 5 families remained indeterminate. Recombination events reduced the 2p minimum candidate region to a 3-cM interval between D2S352 and D2S367, and supported the previously reported 7-cM minimum candidate region for 14q. Age of onset was highly variable, indicating that subtypes of SPG are more appropriately defined on a genetic basis than by age of onset. Comparison of age of onset in parent-child pairs was suggestive of anticipation, with a median difference of 9.0 years (p less than 0.0001).


Molecular Genetics

Hazan et al. (1999) amplified and sequenced overlapping cDNA fragments spanning the entire spastin open reading frame from 1 individual of each of 14 families affected with spastic paraplegia and 6 control individuals. Using this technique, they identified heterozygous mutations in 5 families (604277.0001-604277.0005) with SPG4. Three unrelated affected individuals originating from the same area in Switzerland were heterozygous for a mutation in the acceptor splice site of SPAST intron 15 (604277.0005).

Fonknechten et al. (2000) analyzed DNA from 87 unrelated patients with autosomal dominant hereditary spastic paraplegia and detected 34 novel mutations scattered along the coding region of the SPAST gene (see, e.g., 604277.0007-604277.0008). They found missense (28%), nonsense (15%), and splice site point (26.5%) mutations as well as deletions (23%) and insertions (7.5%). Mean age at onset was 29 +/- 17 years, with a range of 0 to 74 years. Disease severity was highly variable among patients, and disease progression was actually faster in the late-onset group. Penetrance was age-dependent and incomplete even in older mutation carriers (85% after 40 years). Six percent of 238 mutation carriers were asymptomatic, while 20% of carriers were unaware of their symptoms. There was no difference in either age of onset or clinical severity among groups of patients with missense mutations versus truncation mutations.

Svenson et al. (2001) stated that pure hereditary spastic paraplegia type 4 is the most common form of autosomal dominant hereditary SPG. They screened the SPAST gene for mutations in 15 families showing linkage to the SPG4 locus and identified 11 mutations, 10 of which were novel (see, e.g., 604277.0011-604277.0012). In 15 of 76 unrelated individuals with hereditary spastic paraplegia, Meijer et al. (2002) identified 5 previously reported mutations and 8 novel mutations in the SPG4 gene.

Svenson et al. (2004) identified 2 rare polymorphisms in the SPG4 gene: ser44 to leu (S44L; 604277.0015) and pro45 to gln (P45Q; 604277.0017). In affected members of 4 SPG4 families, the presence of either the S44L or P45Q polymorphism in addition to a disease-causing SPG4 mutation (see, e.g., 604277.0016; 604277.0018) resulted in an earlier age at disease onset. Svenson et al. (2004) concluded that the S44L and P45Q polymorphisms, though benign alone, modified the SPG4 phenotype when present with another SPG4 mutation.

Depienne et al. (2006) identified 19 different mutations in the SPG4 gene in 18 (12%) of 146 unrelated mostly European patients with progressive spastic paraplegia. Most of the patients had no family history of the disorder.

In 13 (26%) of 50 unrelated Italian patients with pure hereditary spastic paraplegia (HSP), Crippa et al. (2006) identified 12 different mutations in the SPG4 gene, including 8 novel mutations. All 5 of the familial cases analyzed carried an SPG4 mutation, confirming that the most common form of autosomal dominant HSP is caused by mutations in this gene. Eight (18%) of 45 sporadic patients had a SPG4 mutation. No mutations were identified in 10 additional patients with complicated HSP. Genotype-phenotype correlations were not observed.

In 24 (20%) of 121 probands with autosomal dominant SPG in whom mutations in the SPG4 gene were not detected by DHPLC, Depienne et al. (2007) identified 16 different heterozygous exonic deletions in the SPG4 gene using multiplex ligation-dependent probe amplification (MLPA). The deletions ranged in size from 1 exon to the whole coding sequence. The patients with deletions showed a similar clinical phenotype as those with point mutations but an earlier age at onset. The findings confirmed that haploinsufficiency of SPG4 is a major cause of autosomal dominant SPG and that exonic deletions account for a large proportion of mutation-negative SPG4 patients, justifying the inclusion of gene dosage studies in appropriate clinical scenarios. Depienne et al. (2007) stated that over 150 different pathogenic mutations in the SPG4 gene had been identified to date.

Using MLPA analysis, Beetz et al. (2007) identified partial deletions of the SPG4 gene in 7 of 8 families who had been linked to the region, but in whom mutation screening had not identified mutations. The families had previously been reported by Lindsey et al. (2000), McMonagle et al. (2000), Meijer et al. (2002), and Svenson et al. (2001). The findings indicated that large genomic deletions in SPG4 are not uncommon and should be part of a workup for autosomal dominant SPG.

Mitne-Neto et al. (2007) identified a heterozygous tandem duplication of exons 10 through 12 of the SPG4 gene (604277.0022) in affected individuals of a large Brazilian kindred with spastic paraplegia, originally reported by Starling et al. (2002). In this family, Starling et al. (2002) noted that there were 24 affected men and only 1 affected woman, but X-linked inheritance was ruled out. The authors found strong linkage to the SPG4 locus, but no mutations were identified in the coding region of the SPG4 gene. The results of Mitne-Neto et al. (2007) thus confirmed the diagnosis of SPG4. At the time of the latter report, 12 of 30 mutation carriers had no clinical complaints. Among these patients, 9 of 14 female carriers had no complaints, indicating sex-dependent penetrance in this family, with women being partially protected.

Shoukier et al. (2009) identified SPG4 mutations in 57 (28.5%) of 200 unrelated, mostly German patients with SPG. There were 47 distinct mutations identified, including 29 novel mutations. In a review of other reported mutations, the authors found that most (72.7%) of the mutations were clustered in the C-terminal AAA domain of the SPG4 gene. However, clustering was also observed in the MIT (microtubule interacting and trafficking), MTBD (microtubule-binding domain), and an N-terminal region (residues 228 to 269). In the original cohort of 57 patients, there was a tentative genotype-phenotype correlation indicating that missense mutations were associated with an earlier onset of the disease.


Population Genetics

Among 49 Japanese probands with autosomal dominant SPG who underwent analysis of candidate SPG genes, Ishiura et al. (2014) found that SPG4 was the most common type, accounting for 55.1% of patients.


Animal Model

Du et al. (2010) showed that exogenous expression of wildtype Drosophila or human spastin rescued behavioral and cellular defects in spastin-null flies equivalently. Flies coexpressing 1 copy of wildtype human spastin and 1 copy with the K388R catalytic domain mutation in the fly spastin-null background exhibited aberrant distal synapse morphology and microtubule distribution, similar to but less severe than spastin nulls. R388 or a separate nonsense mutation acted dominantly and were sufficient to confer partial rescue. As in humans, both L44 (604277.0015) and Q45 (604277.0017) were largely silent when heterozygous, but exacerbated mutant phenotypes when expressed in trans with R388.


Cytogenetics

Miura et al. (2011) reported a 4-generation Japanese family from the Miyazaki prefecture in southern Japan with autosomal dominant SPG. RT-PCR and sequencing of affected individuals identified a heterozygous 70-kb deletion of 2p23 encompassing exons 1 to 4 of the SPAST gene as well as exons 1 to 3 of the neighboring DPY30 (612032) gene, located approximately 24 kb upstream of SPAST in a head-to-head orientation. The clinical features included early childhood onset of slowly progressive spastic paraplegia, decreased vibration sense at the ankles, urinary disturbances, and mild cognitive impairment. All 4 affected females had miscarriages, which Miura et al. (2011) speculated may have resulted from loss of DPY30, which plays a role in the regulation of X chromosome dosage compensation and possibly affects sex determination in C. elegans (Hsu and Meyer, 1994).


Nomenclature

Hazan et al. (1993) referred to the form of autosomal dominant spastic paraplegia encoded by a gene on chromosome 14q as FSP1, and Hazan et al. (1994) referred to the form encoded by a gene on chromosome 2p as FSP2. The genes for these 2 disorders are also symbolized SPG3 and SPG4, respectively.


History

Using the repeat expansion detection (RED) method, Nielsen et al. (1997) analyzed 21 affected individuals from 6 SPG4 Danish families with SPG linked to 2p24-p21. They found that 20 of 21 affected individuals showed CAG repeat expansions of the SPG4 gene versus 2 of 21 healthy spouses, demonstrating a strongly statistically significant association between the occurrence of the repeat expansion and the disease. Hazan et al. (1999), however, constructed a detailed high-resolution integrated map of the SPG4 locus that excluded the involvement of a CAG repeat expansion in SPG4-linked autosomal dominant spastic paraplegia. They noted that an analysis of 20 autosomal dominant hereditary spastic paraplegia families, including 4 linked to the SPG4 locus, by Benson et al. (1998) had demonstrated that most repeat expansions detected by the RED method were caused by nonpathogenic expansions at the 18q21.1 SEF2 (602272) and 17q21.3 ERDA1 (603279) loci.


REFERENCES

  1. Beetz, C., Zuchner, S., Ashley-Koch, A., Auer-Grumbach, M., Byrne, P., Chinnery, P. F., Hutchinson, M., McDermott, C. J., Meijer, I. A., Nygren, A. O. H., Pericak-Vance, M., Pyle, A., Rouleau, G. A., Schickel, J., Shaw, P. J., Deufel, T. Linkage to a known gene but no mutation identified: comprehensive reanalysis of SPG4 HSP pedigrees reveals large deletions as the sole cause. (Letter) Hum. Mutat. 28: 739-740, 2007. [PubMed: 17345589, related citations] [Full Text]

  2. Benson, K. F., Horwitz, M., Wolff, J., Friend, K., Thompson, E., White, S., Richards, R. I., Raskind, W. H., Bird, T. D. CAG repeat expansion in autosomal dominant familial spastic paraparesis: novel expansion in a subset of patients. Hum. Molec. Genet. 7: 1779-1786, 1998. [PubMed: 9736780, related citations] [Full Text]

  3. Boustany, R.-M., Fleischnick, E., Alper, C. A., Marazita, M. L., Spence, M. A., Martin, J. B., Kolodny, E. H. The autosomal dominant form of 'pure' familial spastic paraplegia: clinical findings and linkage analysis of a large pedigree. Neurology 37: 910-915, 1987. [PubMed: 3587641, related citations] [Full Text]

  4. Byrne, P. C., Webb, S., McSweeney, F., Burke, T., Hutchinson, M., Parfrey, N. A. Linkage of AD HSP and cognitive impairment to chromosome 2p: haplotype and phenotype analysis indicates variable expression and low or delayed penetrance. Europ. J. Hum. Genet. 6: 275-282, 1998. [PubMed: 9781032, related citations] [Full Text]

  5. Byrne, P., Webb, S., Harbourne, G., MacSweeney, F., Hutchinson, M., Parfrey, N. A. Clinically different forms of hereditary spastic paraplegia map to the same region of chromosome 2. (Abstract) Medizinische Genetik 9: 4 only, 1997.

  6. Crippa, F., Panzeri, C., Martinuzzi, A., Arnoldi, A., Redaelli, F., Tonelli, A., Baschirotto, C., Vazza, G., Mostacciuolo, M. L., Daga, A., Orso, G., Profice, P., and 13 others. Eight novel mutations in SPG4 in a large sample of patients with hereditary spastic paraplegia. Arch. Neurol. 63: 750-755, 2006. [PubMed: 16682546, related citations] [Full Text]

  7. Depienne, C., Fedirko, E., Forlani, S., Cazeneuve, C., Ribai, P., Feki, I., Tallaksen, C., Nguyen, K., Stankoff, B., Ruberg, M., Stevanin, G., Durr, A., Brice, A. Exon deletions of SPG4 are a frequent cause of hereditary spastic paraplegia. (Letter) J. Med. Genet. 44: 281-284, 2007. [PubMed: 17098887, related citations] [Full Text]

  8. Depienne, C., Tallaksen, C., Lephay, J. Y., Bricka, B., Poea-Guyon, S., Fontaine, B., Labauge, P., Brice, A., Durr, A. Spastin mutations are frequent in sporadic spastic paraparesis and their spectrum is different than that observed in familial cases. (Letter) J. Med. Genet. 43: 259-265, 2006. [PubMed: 16055926, related citations] [Full Text]

  9. Du, F., Ozdowski, E. F., Kotowski, I. K., Marchuk, D. A., Sherwood, N. T. Functional conservation of human spastin in a Drosophila model of autosomal dominant-hereditary spastic paraplegia. Hum. Molec. Genet. 19: 1883-1896, 2010. [PubMed: 20154342, images, related citations] [Full Text]

  10. Durr, A., Davoine, C.-S., Paternotte, C., von Fellenberg, J., Cogilinicean, S., Coutinho, P., Lamy, C., Bourgeois, S., Prud'homme, J. F., Penet, C., Burgunder, J. M., Hazan, J., Weissenbach, J., Brice, A., Fontaine, B. Phenotype of autosomal dominant spastic paraplegia linked to chromosome 2. Brain 119: 1487-1496, 1996. [PubMed: 8931574, related citations] [Full Text]

  11. Fonknechten, N., Mavel, D., Byrne, P., Davoine, C.-S., Cruaud, C., Bontsch, D., Samson, D., Coutinho, P., Hutchinson, M., McMonagle, P., Burgunder, J.-M., Tartaglione, A., and 10 others. Spectrum of SPG4 mutations in autosomal dominant spastic paraplegia. Hum. Molec. Genet. 9: 637-644, 2000. Note: Erratum: Hum. Molec. Genet. 14: 461 only, 2005. [PubMed: 10699187, related citations] [Full Text]

  12. Hazan, J., Davoine, C. S., Mavel, D., Fonknechten, N., Paternotte, C., Fizames, C., Cruaud, C., Samson, D., Muselet, D., Vega-Czarny, N., Brice, A., Gyapay, G., Heilig, R., Fontaine, B., Weissenbach, J. A fine integrated map of the SPG4 locus excludes an expanded CAG repeat in chromosome 2p-linked autosomal dominant spastic paraplegia. Genomics 60: 309-319, 1999. [PubMed: 10493830, related citations] [Full Text]

  13. Hazan, J., Fonknechten, N., Mavel, D., Paternotte, C., Samson, D., Artiguenave, F., Davoine, C.-S., Cruaud, C., Durr, A., Wincker, P., Brottier, P., Cattolico, L., Barbe, V., Burgunder, J.-M., Prud'homme, J.-F., Brice, A., Fontaine, B., Heilig, R., Weissenbach, J. Spastin, a new AAA protein, is altered in the most frequent form of autosomal dominant spastic paraplegia. Nature Genet. 23: 296-303, 1999. [PubMed: 10610178, related citations] [Full Text]

  14. Hazan, J., Fontaine, B., Bruyn, R. P. M., Lamy, C., van Deutekom, J. C. T., Rime, C. S., Durr, A., Melki, J., Lyon-Caen, O., Agid, Y., Munnich, A., Padberg, G. W., de Recondo, J., Frants, R. R., Brice, A., Weissenbach, J. Linkage of a new locus for autosomal dominant familial spastic paraplegia to chromosome 2p. Hum. Molec. Genet. 3: 1569-1573, 1994. [PubMed: 7833913, related citations] [Full Text]

  15. Hazan, J., Lamy, C., Melki, J., Munnich, A., de Recondo, J., Weissenbach, J. Autosomal dominant familial spastic paraplegia is genetically heterogeneous and one locus maps to chromosome 14q. Nature Genet. 5: 163-167, 1993. [PubMed: 8252041, related citations] [Full Text]

  16. Hentati, A., Pericak-Vance, M. A., Lennon, F., Wasserman, B., Hentati, F., Juneja, T., Angrist, M. H., Hung, W.-Y., Boustany, R.-M., Bohlega, S., Iqbal, Z., Huether, C. H., Ben Hamida, M., Siddique, T. Linkage of a locus for autosomal dominant familial spastic paraplegia to chromosome 2p markers. Hum. Molec. Genet. 3: 1867-1871, 1994. [PubMed: 7849714, related citations] [Full Text]

  17. Hsu, D. R., Meyer, B. J. The dpy-30 gene encodes an essential component of the Caenorhabditis elegans dosage compensation machinery. Genetics 137: 999-1018, 1994. [PubMed: 7982580, related citations] [Full Text]

  18. Ishiura, H., Takahashi, Y., Hayashi, T., Saito, K., Furuya, H., Watanabe, M., Murata, M., Suzuki, M., Sugiura, A., Sawai, S., Shibuya, K., Ueda, N., Ichikawa, Y., Kanazawa, I., Goto, J., Tsuji, S. Molecular epidemiology and clinical spectrum of hereditary spastic paraplegia in the Japanese population based on comprehensive mutational analyses. J. Hum. Genet. 59: 163-172, 2014. [PubMed: 24451228, related citations] [Full Text]

  19. Lindsey, J. C., Lusher, M. E., McDermott, C. J., White, K. D., Reid, E., Rubinsztein, D. C., Bashir, R., Hazan, J., Shaw, P. J., Bushby, K. M. D. Mutation analysis of the spastin gene (SPG4) in patients with hereditary spastic paraparesis. J. Med. Genet. 37: 759-765, 2000. [PubMed: 11015453, related citations] [Full Text]

  20. McDermott, C. J., Burness, C. E., Kirby, J., Cox, L. E., Rao, D. G., Hewamadduma, C., Sharrack, B., Hadjivassiliou, M., Chinnery, P. F., Dalton, A., Shaw, P. J. Clinical features of hereditary spastic paraplegia due to spastin mutation. Neurology 67: 45-51, 2006. Note: Erratum: Neurology 72: 1534 only, 2009. [PubMed: 16832076, related citations] [Full Text]

  21. McMonagle, P., Byrne, P. C., Fitzgerald, B., Webb, S., Parfrey, N. A., Hutchinson, M. Phenotype of AD-HSP due to mutations in the SPAST gene: comparison with AD-HSP without mutations. Neurology 55: 1794-1800, 2000. [PubMed: 11134375, related citations] [Full Text]

  22. McMonagle, P., Byrne, P., Hutchinson, M. Further evidence of dementia in SPG4-linked autosomal dominant hereditary spastic paraplegia. Neurology 62: 407-410, 2004. [PubMed: 14872021, related citations] [Full Text]

  23. Meijer, I. A., Hand, C. K., Cossette, P., Figlewicz, D. A., Rouleau, G. A. Spectrum of SPG4 mutations in a large collection of North American families with hereditary spastic paraplegia. Arch. Neurol. 59: 281-286, 2002. [PubMed: 11843700, related citations] [Full Text]

  24. Mitne-Neto, M., Kok, F., Beetz, C., Pessoa, A., Bueno, C., Graciani, Z., Martyn, M., Monteiro, C. B. M., Mitne, G., Hubert, P., Nygren, A. O. H., Valadares, M., Cerqueira, A. M. P., Starling, A., Deufel, T., Zatz, M. A multi-exonic SPG4 duplication underlies sex-dependent penetrance of hereditary spastic paraplegia in a large Brazilian pedigree. Europ. J. Hum. Genet. 15: 1276-1279, 2007. [PubMed: 17895902, related citations] [Full Text]

  25. Miura, S., Shibata, H., Kida, H., Noda, K., Toyama, T., Iwasaki, N., Iwaki, A., Ayabe, M., Aizawa, H., Taniwaki, T., Fukumaki, Y. Partial SPAST and DPY30 deletions in a Japanese spastic paraplegia type 4 family. Neurogenetics 12: 25-31, 2011. [PubMed: 20857310, related citations] [Full Text]

  26. Murphy, S., Gorman, G., Beetz, C., Byrne, P., Dytko, M., McMonagle, P., Kinsella, K., Farrell, M., Hutchinson, M. Dementia in SPG4 hereditary spastic paraplegia: clinical, genetic, and neuropathologic evidence. Neurology 73: 378-384, 2009. [PubMed: 19652142, related citations] [Full Text]

  27. Nance, M. A., Raabe, W. A., Midani, H., Kolodny, E. H., David, W. S., Megna, L., Pericak-Vance, M. A., Haines, J. L. Clinical heterogeneity of familial spastic paraplegia linked to chromosome 2p21. Hum. Hered. 48: 169-178, 1998. [PubMed: 9618065, related citations] [Full Text]

  28. Nielsen, J. E., Koefoed, P., Abell, K., Hasholt, L., Eiberg, H., Fenger, K., Niebuhr, E., Sorensen, S. A. CAG repeat expansion in autosomal dominant pure spastic paraplegia linked to chromosome 2p21-p24. Hum. Molec. Genet. 6: 1811-1816, 1997. [PubMed: 9302257, related citations] [Full Text]

  29. Nielsen, J. E., Krabbe, K., Jennum, P., Koefoed, P., Jensen, L. N., Fenger, K., Eiberg, H., Hasholt, L., Werdelin, L., Sorensen, S. A. Autosomal dominant pure spastic paraplegia: a clinical, paraclinical, and genetic study. J. Neurol. Neurosurg. Psychiat. 64: 61-66, 1998. [PubMed: 9436729, related citations] [Full Text]

  30. Orlacchio, A., Gaudiello, F., Totaro, A., Floris, R., St George-Hyslop, P. H., Bernardi, G., Kawarai, T. A new SPG4 mutation in a variant form of spastic paraplegia with congenital arachnoid cysts. Neurology 62: 1875-1878, 2004. [PubMed: 15159500, related citations] [Full Text]

  31. Orlacchio, A., Kawarai, T., Totaro, A., Errico, A., St George-Hyslop, P. H., Rugarli, E. I., Bernardi, G. Hereditary spastic paraplegia: clinical genetic study of 15 families. Arch. Neurol. 61: 849-855, 2004. [PubMed: 15210521, related citations] [Full Text]

  32. Orlacchio, A., Patrono, C., Gaudiello, F., Rocchi, C., Moschella, V., Floris, R., Bernardi, G., Kawarai, T. Silver syndrome variant of hereditary spastic paraplegia: a locus to 4p and allelism with SPG4. Neurology 70: 1959-1966, 2008. [PubMed: 18401025, related citations] [Full Text]

  33. Reid, E., Grayson, C., Rubinsztein, D. C., Rogers, M. T., Rubinsztein, J. S. Subclinical cognitive impairment in autosomal dominant 'pure' hereditary spastic paraplegia. J. Med. Genet. 36: 797-798, 1999. [PubMed: 10528866, related citations] [Full Text]

  34. Sack, G. H., Jr., Huether, C. A., Garg, N. Familial spastic paraplegia--clinical and pathologic studies in a large kindred. Johns Hopkins Med. J. 143: 117-121, 1978. [PubMed: 703033, related citations]

  35. Scott, W. K., Gaskell, P. C., Lennon, F., Wolpert, C. M., Menold, M. M., Aylsworth, A. S., Warner, C., Farrell, C. D., Boustany, R.-M. N., Albright, S. G., Boyd, E., Kingston, H. M., Cumming, W. J. K., Vance, J. M., Pericak-Vance, M. A. Locus heterogeneity, anticipation and reduction of the chromosome 2p minimal candidate region in autosomal dominant familial spastic paraplegia. Neurogenetics 1: 95-102, 1997. [PubMed: 10732810, related citations] [Full Text]

  36. Shoukier, M., Neesen, J., Sauter, S. M., Argyriou, L., Doerwald, N., Pantakani, D. V. K., Mannan, A. U. Expansion of mutation spectrum, determination of mutation cluster regions and predictive structural classification of SPAST mutations in hereditary spastic paraplegia. Europ. J. Hum. Genet. 17: 187-194, 2009. Note: Erratum: Europ. J. Hum. Genet. 17: 401-402, 2009. [PubMed: 18701882, related citations] [Full Text]

  37. Starling, A., Rocco, P., Passos-Bueno, M. R., Hazan, J., Marie, S. K., Zatz, M. Autosomal dominant (AD) pure spastic paraplegia (HSP) linked to locus SPG4 affects almost exclusively males in a large pedigree. J. Med. Genet. 39: e77, 2002. Note: Electronic article. [PubMed: 12471215, related citations] [Full Text]

  38. Svenson, I. K., Ashley-Koch, A. E., Gaskell, P. C., Riney, T. J., Cumming, W. J. K., Kingston, H. M., Hogan, E. L., Boustany, R.-M. N., Vance, J. M., Nance, M. A., Pericak-Vance, M. A., Marchuk, D. A. Identification and expression analysis of spastin gene mutations in hereditary spastic paraplegia. Am. J. Hum. Genet. 68: 1077-1085, 2001. [PubMed: 11309678, images, related citations] [Full Text]

  39. Svenson, I. K., Kloos, M. T., Gaskell, P. C., Nance, M. A., Garbern, J. Y., Hisanaga, S., Pericak-Vance, M. A., Ashley-Koch, A. E., Marchuk, D. A. Intragenic modifiers of hereditary spastic paraplegia due to spastin gene mutations. Neurogenetics 5: 157-164, 2004. [PubMed: 15248095, related citations] [Full Text]

  40. White, K. D., Ince, P. G., Lusher, M., Lindsey, J., Cookson, M., Bashir, R., Shaw, P. J., Bushby, K. M. D. Clinical and pathologic findings in hereditary spastic paraparesis with spastin mutation. Neurology 55: 89-94, 2000. [PubMed: 10891911, related citations] [Full Text]


Cassandra L. Kniffin - updated : 4/21/2014
Cassandra L. Kniffin - updated : 2/12/2013
Cassandra L. Kniffin - updated : 12/17/2009
Cassandra L. Kniffin - updated : 8/28/2009
Cassandra L. Kniffin - updated : 10/1/2008
Cassandra L. Kniffin - updated : 3/19/2008
Cassandra L. Kniffin - updated : 8/20/2007
Cassandra L. Kniffin - updated : 7/24/2007
Cassandra L. Kniffin - updated : 4/27/2007
Cassandra L. Kniffin - updated : 2/6/2007
Cassandra L. Kniffin - updated : 4/11/2006
Cassandra L. Kniffin - updated : 1/26/2005
Cassandra L. Kniffin - updated : 10/26/2004
Cassandra L. Kniffin - updated : 8/30/2004
Cassandra L. Kniffin - updated : 7/27/2004
Cassandra L. Kniffin - updated : 12/27/2002
Cassandra L. Kniffin - reorganized : 10/4/2002
Cassandra L. Kniffin - updated : 6/26/2002
Victor A. McKusick - updated : 6/15/2001
George E. Tiller - updated : 4/14/2000
Michael J. Wright - updated : 1/19/2000
Victor A. McKusick - updated : 11/2/1998
Victor A. McKusick - updated : 10/2/1998
Victor A. McKusick - updated : 8/17/1998
Victor A. McKusick - updated : 7/7/1998
Victor A. McKusick - updated : 5/5/1998
Victor A. McKusick - updated : 11/4/1997
Victor A. McKusick - updated : 5/30/1997
Creation Date:
Victor A. McKusick : 10/14/1993
alopez : 11/17/2023
alopez : 11/17/2023
carol : 04/10/2023
carol : 10/09/2019
carol : 10/08/2019
carol : 08/31/2016
carol : 04/23/2014
ckniffin : 4/21/2014
tpirozzi : 8/14/2013
tpirozzi : 8/14/2013
tpirozzi : 8/14/2013
tpirozzi : 8/13/2013
terry : 3/15/2013
carol : 3/11/2013
carol : 3/11/2013
alopez : 2/18/2013
ckniffin : 2/12/2013
wwang : 3/2/2011
wwang : 1/8/2010
ckniffin : 12/17/2009
wwang : 9/14/2009
ckniffin : 8/28/2009
wwang : 11/11/2008
wwang : 10/3/2008
ckniffin : 10/1/2008
wwang : 4/10/2008
ckniffin : 3/19/2008
wwang : 9/6/2007
ckniffin : 8/20/2007
wwang : 8/3/2007
ckniffin : 7/24/2007
wwang : 6/8/2007
ckniffin : 4/27/2007
wwang : 2/8/2007
ckniffin : 2/6/2007
wwang : 4/20/2006
ckniffin : 4/11/2006
terry : 6/24/2005
tkritzer : 2/2/2005
ckniffin : 1/26/2005
tkritzer : 11/1/2004
ckniffin : 10/26/2004
carol : 9/7/2004
ckniffin : 8/30/2004
tkritzer : 7/28/2004
ckniffin : 7/27/2004
carol : 1/6/2003
ckniffin : 12/27/2002
carol : 10/4/2002
ckniffin : 9/30/2002
tkritzer : 8/9/2002
ckniffin : 6/26/2002
cwells : 6/27/2001
terry : 6/15/2001
alopez : 4/17/2000
terry : 4/14/2000
alopez : 1/19/2000
alopez : 11/9/1999
alopez : 11/2/1999
mgross : 9/24/1999
carol : 11/11/1998
terry : 11/2/1998
dkim : 10/12/1998
carol : 10/7/1998
terry : 10/2/1998
carol : 8/18/1998
terry : 8/17/1998
carol : 7/9/1998
terry : 7/7/1998
carol : 5/12/1998
terry : 5/5/1998
jenny : 11/12/1997
terry : 11/4/1997
alopez : 7/29/1997
terry : 7/7/1997
jenny : 6/3/1997
terry : 5/30/1997
mimadm : 3/25/1995
carol : 1/23/1995
terry : 11/16/1994
carol : 10/14/1993

# 182601

SPASTIC PARAPLEGIA 4, AUTOSOMAL DOMINANT; SPG4


Alternative titles; symbols

FAMILIAL SPASTIC PARAPLEGIA, AUTOSOMAL DOMINANT, 2; FSP2


SNOMEDCT: 723820001;   ORPHA: 100985;   DO: 0110792;  


Phenotype-Gene Relationships

Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
Gene/Locus Gene/Locus
MIM number
2p22.3 Spastic paraplegia 4, autosomal dominant 182601 Autosomal dominant 3 SPAST 604277

TEXT

A number sign (#) is used with this entry because autosomal dominant spastic paraplegia-4 (SPG4) is caused by heterozygous mutation in the SPAST gene (604277) on chromosome 2p22.


Description

The hereditary spastic paraplegias (SPG, HSP) are a group of clinically and genetically diverse inherited disorders characterized predominantly by progressive lower extremity spasticity and weakness. SPG is classified by mode of inheritance (autosomal dominant, autosomal recessive, and X-linked) and whether the primary symptoms occur in isolation ('uncomplicated') or with other neurologic abnormalities ('complicated').

Pure SPG4 is the most common form of autosomal dominant hereditary SPG, comprising up to 45% of cases (Svenson et al., 2001; Crippa et al., 2006).

For a general phenotypic description and a discussion of genetic heterogeneity of autosomal dominant spastic paraplegia, see SPG3A (182600).


Clinical Features

In 5 of 7 French families and in 1 large Dutch pedigree with a form of autosomal dominant familial spastic paraplegia, Hazan et al. (1994) found linkage to a locus, which they termed FSP2 (also known as SPG4), on chromosome 2p. This finding distinguished the disease from autosomal dominant spastic paraplegia-3 (182600), which had been mapped to chromosome 14. Age of onset in the 6 families showing linkage to 2p varied widely within families and the mean age at onset ranged from 20 to 39 years. Thus, age of onset may be a poor criterion for classifying autosomal dominant spastic paraplegia. Anticipation in the age of onset was observed in 2 of the kindreds.

Durr et al. (1996) reported 12 families with autosomal dominant spastic paraplegia linked to the SPG4 locus on chromosome 2. Age of onset ranged from infancy to 63 years. The clinical expression of the disorder within a family included asymptomatic patients who were unaware of their condition, mildly affected individuals who had spastic gait but were able to walk independently, and severely affected patients who were wheelchair bound. Durr et al. (1996) commented on the extensive intra- and interfamilial clinical variation.

Nielsen et al. (1998) evaluated 5 families with 2p-linked pure spastic paraplegia. In 2 families, nonprogressive 'congenital' spastic paraplegia was seen in some affected members, whereas adult-onset progressive spastic paraplegia was present in others. Low backache was reported as a late symptom by 47% of the 63 at-risk members in the 5 families. Brain and total spinal cord MRI disclosed no significant abnormalities. Nielsen et al. (1998) concluded that SPG4 is a phenotypically heterogeneous disorder, characterized by both interfamilial and intrafamilial variation.

Nance et al. (1998) found striking variation in clinical features in 4 families with spastic paraplegia with linkage to chromosome 2 markers. Only mild neurologic signs were observed in some subjects. The clinical features of 1 family had previously been described by Boustany et al. (1987). Onset was generally in the third to fifth decades with an average onset age of 35 years (range, 5 to 61 years). All clearly affected patients had scissoring gait, and in all who were examined at least 2 of the following were found: extensor plantar responses, increased knee and ankle reflexes, increased tone, muscle spasms, or leg cramps. Urinary urgency or other symptoms compatible with a neurogenic bladder, leg weakness, and decreased vibration sense were present in some, but not all, patients.

Byrne et al. (1997, 1998) presented a family with autosomal dominant hereditary spastic paraplegia and a specific form of cognitive impairment who showed linkage to the SPG4 locus on chromosome 2. The pattern of cognitive impairment in this family was characterized primarily by deficits in visual-spatial functions. Dysfunction manifested itself by difficulty in carrying out new tasks, forgetfulness, poor spatial perception, and impaired visual-motor coordination. By haplotype analysis the presence of the mutant gene was identified in an individual who, at the age of 57, had the same pattern of cognitive impairment but no spastic paraplegia. Furthermore, 6 individuals who presented with the disease haplotype had normal neurologic and neuropsychologic examinations. All 6 were below the maximal age of onset in the family, namely, 60 years. In this Irish family the cognitive impairment was considered to be a manifestation of the SPG4 gene mutation.

Reid et al. (1999) investigated 35 individuals from 4 families of Welsh origin, 22 of whom had 'pure' hereditary spastic paraplegia, for the presence of subclinical cognitive impairment. They found significant reductions in scores on the Mini-Mental State Examination (MMSE) among affected individuals compared to controls. To assess whether the lower MMSE scores were restricted to subjects older than 50 years, scores for affected subjects 50 years of age or younger were compared to those of controls. A significant difference in score remained. One of the families was linked to the chromosome 2 locus, while 2 others showed linkage to none of the loci known at that time. There was no significant difference between the results of these 2 groups.

McMonagle et al. (2000) compared the phenotypic expressions of autosomal dominant hereditary spastic paresis in several families with a mutation in the SPG4 gene and several families without a mutation in SPG4. In the mutation-positive group, age of onset was later, disability score was greater, progression of disease was faster, wheelchair use was greater (40.9% vs 4.8% in the mutation-excluded group), there was greater abnormal vibration sensation in the lower limbs (68.2% vs 19%), and fewer individuals were asymptomatic (18.2% vs 42.9%). Dementia was more prevalent in the mutation-positive group. McMonagle et al. (2000) emphasized the finding of cognitive impairment as a feature of SPG4 mutations.

White et al. (2000) reported a patient with familial SPG4 who had clinical dementia. Postmortem neuropathologic examination showed neuronal loss and tau- (MAPT; 157140) immunoreactive neurofibrillary tangles in the hippocampus and tau-immunoreactive balloon cells in the limbic area and neocortex. Lewy bodies were present in the substantia nigra. White et al. (2000) suggested that these findings confirmed an association of dementia with SPG4.

McMonagle et al. (2004) used several measures of cognitive function to assess 11 patients from 3 families in whom SPG4 was confirmed by genetic analysis or linkage. SPG4 patients scored significantly lower on the Cambridge cognitive examination (CAMCOG) (mean score of 73.5 compared to 91.7 in controls). After approximately 3 years, the patients' mean score fell to 64.4, whereas the mean control score declined slightly to 90.8. Deficits in the SPG4 patients were noted in attention, language expression, memory, and abstraction. Behavior assessment found that SPG4 patients exhibited agitation, aggression, apathy, irritability, depression, and disinhibition. Accounting for age, McMonagle et al. (2004) concluded that subtle changes in cognitive function in patients with SPG4 may begin after age 40 years, with more severe decline after age 60.

Orlacchio et al. (2004) reported 32 patients from 9 families from southern Scotland with SPG4. Age at onset varied from 11 to 53 years. In addition to classic features of hereditary spastic paraplegia, 2 of the 32 patients had mental retardation and 2 other patients had a thin corpus callosum and cerebellar atrophy. All affected members had the same mutation in the SPG4 gene (604277.0014), and haplotype analysis suggested a founder effect.

Orlacchio et al. (2004) reported a large Italian family in which all 16 members who had SPG4 also had congenital arachnoid cysts at the cerebellopontine angle ranging in size from 21 to 31 mm. Six patients also had mental retardation. Genetic analysis confirmed a mutation in the SPG4 gene.

McDermott et al. (2006) reported a patient with SPG4 who developed walking difficulties in his late teens with deteriorating gait in his 20s; he was wheelchair-dependent at age 35. He later developed stiffness in the upper limbs, bladder dysfunction, dysarthria, and swallowing difficulties. In his 40s, he developed respiratory insufficiency and distal muscle wasting in the lower limbs. Molecular analysis identified a mutation in the SPG4 gene (S445R; 604277.0021). The findings of bulbar and respiratory involvement, as well as lower motor neuron degeneration, broadened the phenotype associated with mutations in the SPG4 gene.

Orlacchio et al. (2008) reported a large 4-generation Italian family with SPG4 confirmed by genetic analysis. The mean ages at onset were 17.5 and 18.8 years for symptoms of the lower and upper limbs, respectively. There was a general impression of genetic anticipation spanning the 4 generations. All affected individuals had spasticity of the lower limbs and pyramidal tract signs such as hyperreflexia, extensor plantar responses, or both, and pes cavus. All patients also had weak intrinsic hand muscles, with severe amyotrophy most relevant in the thenar eminence. Peroneal muscle wasting was reported in five patients, and many used a cane. Other associated features included impaired vibration sensation and cognitive dysfunctions. All patients except 1 had temporal lobe epilepsy with partial complex seizures associated with hippocampal sclerosis.

Murphy et al. (2009) reported a family in which 12 members had SPG4 due to a deletion of exon 17 in the SPG4 gene (Beetz et al., 2007). Cognitive assessment performed over a 7-year period found that all 4 patients who were older than 60 years developed mild to moderate cognitive decline. Two younger patients aged 48 and 40, respectively, had mild cognitive impairment. Genetic analysis of this family was unusual because 4 patients with the SPG4 deletion also carried a microdeletion in the NIPA1 gene (608145), which causes SPG6 (600363); only 2 of these 4 had cognitive impairment. Five patients with only the SPG4 deletion had cognitive impairment, including 2 who did not have clinical signs of SPG. Another family member with only the NIPA1 microdeletion lacked clinical signs of SPG or cognitive impairment at age 57. Murphy et al. (2009) concluded that SPG4 is associated with cognitive decline, and that the SPG6 microdeletion does not have a clinical phenotype in this family. Postmortem examination of the proband, who had both deletions as well as SPG and cognitive impairment, showed a markedly atrophic spinal cord with degeneration of the corticospinal tracts, and superficial spongiosis and widespread ubiquitin-positive inclusions in the neocortex and white matter.


Inheritance

The transmission pattern of SPG4 in the families reported by Hazan et al. (1999) was consistent with autosomal dominant inheritance.


Mapping

In 5 French families and 1 large Dutch pedigree with autosomal dominant spastic paraplegia, Hazan et al. (1994) found linkage markers in the 2p24-p21 region. An analysis of recombination events and multipoint linkage placed this form of the disease within a 4-cM interval flanked by loci D2S400 and D2S367. In 4 Caucasian North American families and in 1 family from Tunisia, Hentati et al. (1994) found linkage of late-onset SPG to DNA markers on chromosome 2p in 4 of the families. Pathologic findings in a member of one of the chromosome 2-linked families had previously been reported by Sack et al. (1978).

Scott et al. (1997) examined 11 Caucasian pedigrees with autosomal dominant 'uncomplicated' familial spastic paraplegia for linkage to the previously identified loci on 2p, 14q (SPG3A), and 15q (SPG6; 600363). Chromosome 15q was excluded for all families. Five families showed evidence for linkage to 2p, 1 family to 14q, and 5 families remained indeterminate. Recombination events reduced the 2p minimum candidate region to a 3-cM interval between D2S352 and D2S367, and supported the previously reported 7-cM minimum candidate region for 14q. Age of onset was highly variable, indicating that subtypes of SPG are more appropriately defined on a genetic basis than by age of onset. Comparison of age of onset in parent-child pairs was suggestive of anticipation, with a median difference of 9.0 years (p less than 0.0001).


Molecular Genetics

Hazan et al. (1999) amplified and sequenced overlapping cDNA fragments spanning the entire spastin open reading frame from 1 individual of each of 14 families affected with spastic paraplegia and 6 control individuals. Using this technique, they identified heterozygous mutations in 5 families (604277.0001-604277.0005) with SPG4. Three unrelated affected individuals originating from the same area in Switzerland were heterozygous for a mutation in the acceptor splice site of SPAST intron 15 (604277.0005).

Fonknechten et al. (2000) analyzed DNA from 87 unrelated patients with autosomal dominant hereditary spastic paraplegia and detected 34 novel mutations scattered along the coding region of the SPAST gene (see, e.g., 604277.0007-604277.0008). They found missense (28%), nonsense (15%), and splice site point (26.5%) mutations as well as deletions (23%) and insertions (7.5%). Mean age at onset was 29 +/- 17 years, with a range of 0 to 74 years. Disease severity was highly variable among patients, and disease progression was actually faster in the late-onset group. Penetrance was age-dependent and incomplete even in older mutation carriers (85% after 40 years). Six percent of 238 mutation carriers were asymptomatic, while 20% of carriers were unaware of their symptoms. There was no difference in either age of onset or clinical severity among groups of patients with missense mutations versus truncation mutations.

Svenson et al. (2001) stated that pure hereditary spastic paraplegia type 4 is the most common form of autosomal dominant hereditary SPG. They screened the SPAST gene for mutations in 15 families showing linkage to the SPG4 locus and identified 11 mutations, 10 of which were novel (see, e.g., 604277.0011-604277.0012). In 15 of 76 unrelated individuals with hereditary spastic paraplegia, Meijer et al. (2002) identified 5 previously reported mutations and 8 novel mutations in the SPG4 gene.

Svenson et al. (2004) identified 2 rare polymorphisms in the SPG4 gene: ser44 to leu (S44L; 604277.0015) and pro45 to gln (P45Q; 604277.0017). In affected members of 4 SPG4 families, the presence of either the S44L or P45Q polymorphism in addition to a disease-causing SPG4 mutation (see, e.g., 604277.0016; 604277.0018) resulted in an earlier age at disease onset. Svenson et al. (2004) concluded that the S44L and P45Q polymorphisms, though benign alone, modified the SPG4 phenotype when present with another SPG4 mutation.

Depienne et al. (2006) identified 19 different mutations in the SPG4 gene in 18 (12%) of 146 unrelated mostly European patients with progressive spastic paraplegia. Most of the patients had no family history of the disorder.

In 13 (26%) of 50 unrelated Italian patients with pure hereditary spastic paraplegia (HSP), Crippa et al. (2006) identified 12 different mutations in the SPG4 gene, including 8 novel mutations. All 5 of the familial cases analyzed carried an SPG4 mutation, confirming that the most common form of autosomal dominant HSP is caused by mutations in this gene. Eight (18%) of 45 sporadic patients had a SPG4 mutation. No mutations were identified in 10 additional patients with complicated HSP. Genotype-phenotype correlations were not observed.

In 24 (20%) of 121 probands with autosomal dominant SPG in whom mutations in the SPG4 gene were not detected by DHPLC, Depienne et al. (2007) identified 16 different heterozygous exonic deletions in the SPG4 gene using multiplex ligation-dependent probe amplification (MLPA). The deletions ranged in size from 1 exon to the whole coding sequence. The patients with deletions showed a similar clinical phenotype as those with point mutations but an earlier age at onset. The findings confirmed that haploinsufficiency of SPG4 is a major cause of autosomal dominant SPG and that exonic deletions account for a large proportion of mutation-negative SPG4 patients, justifying the inclusion of gene dosage studies in appropriate clinical scenarios. Depienne et al. (2007) stated that over 150 different pathogenic mutations in the SPG4 gene had been identified to date.

Using MLPA analysis, Beetz et al. (2007) identified partial deletions of the SPG4 gene in 7 of 8 families who had been linked to the region, but in whom mutation screening had not identified mutations. The families had previously been reported by Lindsey et al. (2000), McMonagle et al. (2000), Meijer et al. (2002), and Svenson et al. (2001). The findings indicated that large genomic deletions in SPG4 are not uncommon and should be part of a workup for autosomal dominant SPG.

Mitne-Neto et al. (2007) identified a heterozygous tandem duplication of exons 10 through 12 of the SPG4 gene (604277.0022) in affected individuals of a large Brazilian kindred with spastic paraplegia, originally reported by Starling et al. (2002). In this family, Starling et al. (2002) noted that there were 24 affected men and only 1 affected woman, but X-linked inheritance was ruled out. The authors found strong linkage to the SPG4 locus, but no mutations were identified in the coding region of the SPG4 gene. The results of Mitne-Neto et al. (2007) thus confirmed the diagnosis of SPG4. At the time of the latter report, 12 of 30 mutation carriers had no clinical complaints. Among these patients, 9 of 14 female carriers had no complaints, indicating sex-dependent penetrance in this family, with women being partially protected.

Shoukier et al. (2009) identified SPG4 mutations in 57 (28.5%) of 200 unrelated, mostly German patients with SPG. There were 47 distinct mutations identified, including 29 novel mutations. In a review of other reported mutations, the authors found that most (72.7%) of the mutations were clustered in the C-terminal AAA domain of the SPG4 gene. However, clustering was also observed in the MIT (microtubule interacting and trafficking), MTBD (microtubule-binding domain), and an N-terminal region (residues 228 to 269). In the original cohort of 57 patients, there was a tentative genotype-phenotype correlation indicating that missense mutations were associated with an earlier onset of the disease.


Population Genetics

Among 49 Japanese probands with autosomal dominant SPG who underwent analysis of candidate SPG genes, Ishiura et al. (2014) found that SPG4 was the most common type, accounting for 55.1% of patients.


Animal Model

Du et al. (2010) showed that exogenous expression of wildtype Drosophila or human spastin rescued behavioral and cellular defects in spastin-null flies equivalently. Flies coexpressing 1 copy of wildtype human spastin and 1 copy with the K388R catalytic domain mutation in the fly spastin-null background exhibited aberrant distal synapse morphology and microtubule distribution, similar to but less severe than spastin nulls. R388 or a separate nonsense mutation acted dominantly and were sufficient to confer partial rescue. As in humans, both L44 (604277.0015) and Q45 (604277.0017) were largely silent when heterozygous, but exacerbated mutant phenotypes when expressed in trans with R388.


Cytogenetics

Miura et al. (2011) reported a 4-generation Japanese family from the Miyazaki prefecture in southern Japan with autosomal dominant SPG. RT-PCR and sequencing of affected individuals identified a heterozygous 70-kb deletion of 2p23 encompassing exons 1 to 4 of the SPAST gene as well as exons 1 to 3 of the neighboring DPY30 (612032) gene, located approximately 24 kb upstream of SPAST in a head-to-head orientation. The clinical features included early childhood onset of slowly progressive spastic paraplegia, decreased vibration sense at the ankles, urinary disturbances, and mild cognitive impairment. All 4 affected females had miscarriages, which Miura et al. (2011) speculated may have resulted from loss of DPY30, which plays a role in the regulation of X chromosome dosage compensation and possibly affects sex determination in C. elegans (Hsu and Meyer, 1994).


Nomenclature

Hazan et al. (1993) referred to the form of autosomal dominant spastic paraplegia encoded by a gene on chromosome 14q as FSP1, and Hazan et al. (1994) referred to the form encoded by a gene on chromosome 2p as FSP2. The genes for these 2 disorders are also symbolized SPG3 and SPG4, respectively.


History

Using the repeat expansion detection (RED) method, Nielsen et al. (1997) analyzed 21 affected individuals from 6 SPG4 Danish families with SPG linked to 2p24-p21. They found that 20 of 21 affected individuals showed CAG repeat expansions of the SPG4 gene versus 2 of 21 healthy spouses, demonstrating a strongly statistically significant association between the occurrence of the repeat expansion and the disease. Hazan et al. (1999), however, constructed a detailed high-resolution integrated map of the SPG4 locus that excluded the involvement of a CAG repeat expansion in SPG4-linked autosomal dominant spastic paraplegia. They noted that an analysis of 20 autosomal dominant hereditary spastic paraplegia families, including 4 linked to the SPG4 locus, by Benson et al. (1998) had demonstrated that most repeat expansions detected by the RED method were caused by nonpathogenic expansions at the 18q21.1 SEF2 (602272) and 17q21.3 ERDA1 (603279) loci.


REFERENCES

  1. Beetz, C., Zuchner, S., Ashley-Koch, A., Auer-Grumbach, M., Byrne, P., Chinnery, P. F., Hutchinson, M., McDermott, C. J., Meijer, I. A., Nygren, A. O. H., Pericak-Vance, M., Pyle, A., Rouleau, G. A., Schickel, J., Shaw, P. J., Deufel, T. Linkage to a known gene but no mutation identified: comprehensive reanalysis of SPG4 HSP pedigrees reveals large deletions as the sole cause. (Letter) Hum. Mutat. 28: 739-740, 2007. [PubMed: 17345589] [Full Text: https://doi.org/10.1002/humu.20508]

  2. Benson, K. F., Horwitz, M., Wolff, J., Friend, K., Thompson, E., White, S., Richards, R. I., Raskind, W. H., Bird, T. D. CAG repeat expansion in autosomal dominant familial spastic paraparesis: novel expansion in a subset of patients. Hum. Molec. Genet. 7: 1779-1786, 1998. [PubMed: 9736780] [Full Text: https://doi.org/10.1093/hmg/7.11.1779]

  3. Boustany, R.-M., Fleischnick, E., Alper, C. A., Marazita, M. L., Spence, M. A., Martin, J. B., Kolodny, E. H. The autosomal dominant form of 'pure' familial spastic paraplegia: clinical findings and linkage analysis of a large pedigree. Neurology 37: 910-915, 1987. [PubMed: 3587641] [Full Text: https://doi.org/10.1212/wnl.37.6.910]

  4. Byrne, P. C., Webb, S., McSweeney, F., Burke, T., Hutchinson, M., Parfrey, N. A. Linkage of AD HSP and cognitive impairment to chromosome 2p: haplotype and phenotype analysis indicates variable expression and low or delayed penetrance. Europ. J. Hum. Genet. 6: 275-282, 1998. [PubMed: 9781032] [Full Text: https://doi.org/10.1038/sj.ejhg.5200185]

  5. Byrne, P., Webb, S., Harbourne, G., MacSweeney, F., Hutchinson, M., Parfrey, N. A. Clinically different forms of hereditary spastic paraplegia map to the same region of chromosome 2. (Abstract) Medizinische Genetik 9: 4 only, 1997.

  6. Crippa, F., Panzeri, C., Martinuzzi, A., Arnoldi, A., Redaelli, F., Tonelli, A., Baschirotto, C., Vazza, G., Mostacciuolo, M. L., Daga, A., Orso, G., Profice, P., and 13 others. Eight novel mutations in SPG4 in a large sample of patients with hereditary spastic paraplegia. Arch. Neurol. 63: 750-755, 2006. [PubMed: 16682546] [Full Text: https://doi.org/10.1001/archneur.63.5.750]

  7. Depienne, C., Fedirko, E., Forlani, S., Cazeneuve, C., Ribai, P., Feki, I., Tallaksen, C., Nguyen, K., Stankoff, B., Ruberg, M., Stevanin, G., Durr, A., Brice, A. Exon deletions of SPG4 are a frequent cause of hereditary spastic paraplegia. (Letter) J. Med. Genet. 44: 281-284, 2007. [PubMed: 17098887] [Full Text: https://doi.org/10.1136/jmg.2006.046425]

  8. Depienne, C., Tallaksen, C., Lephay, J. Y., Bricka, B., Poea-Guyon, S., Fontaine, B., Labauge, P., Brice, A., Durr, A. Spastin mutations are frequent in sporadic spastic paraparesis and their spectrum is different than that observed in familial cases. (Letter) J. Med. Genet. 43: 259-265, 2006. [PubMed: 16055926] [Full Text: https://doi.org/10.1136/jmg.2005.035311]

  9. Du, F., Ozdowski, E. F., Kotowski, I. K., Marchuk, D. A., Sherwood, N. T. Functional conservation of human spastin in a Drosophila model of autosomal dominant-hereditary spastic paraplegia. Hum. Molec. Genet. 19: 1883-1896, 2010. [PubMed: 20154342] [Full Text: https://doi.org/10.1093/hmg/ddq064]

  10. Durr, A., Davoine, C.-S., Paternotte, C., von Fellenberg, J., Cogilinicean, S., Coutinho, P., Lamy, C., Bourgeois, S., Prud'homme, J. F., Penet, C., Burgunder, J. M., Hazan, J., Weissenbach, J., Brice, A., Fontaine, B. Phenotype of autosomal dominant spastic paraplegia linked to chromosome 2. Brain 119: 1487-1496, 1996. [PubMed: 8931574] [Full Text: https://doi.org/10.1093/brain/119.5.1487]

  11. Fonknechten, N., Mavel, D., Byrne, P., Davoine, C.-S., Cruaud, C., Bontsch, D., Samson, D., Coutinho, P., Hutchinson, M., McMonagle, P., Burgunder, J.-M., Tartaglione, A., and 10 others. Spectrum of SPG4 mutations in autosomal dominant spastic paraplegia. Hum. Molec. Genet. 9: 637-644, 2000. Note: Erratum: Hum. Molec. Genet. 14: 461 only, 2005. [PubMed: 10699187] [Full Text: https://doi.org/10.1093/hmg/9.4.637]

  12. Hazan, J., Davoine, C. S., Mavel, D., Fonknechten, N., Paternotte, C., Fizames, C., Cruaud, C., Samson, D., Muselet, D., Vega-Czarny, N., Brice, A., Gyapay, G., Heilig, R., Fontaine, B., Weissenbach, J. A fine integrated map of the SPG4 locus excludes an expanded CAG repeat in chromosome 2p-linked autosomal dominant spastic paraplegia. Genomics 60: 309-319, 1999. [PubMed: 10493830] [Full Text: https://doi.org/10.1006/geno.1999.5932]

  13. Hazan, J., Fonknechten, N., Mavel, D., Paternotte, C., Samson, D., Artiguenave, F., Davoine, C.-S., Cruaud, C., Durr, A., Wincker, P., Brottier, P., Cattolico, L., Barbe, V., Burgunder, J.-M., Prud'homme, J.-F., Brice, A., Fontaine, B., Heilig, R., Weissenbach, J. Spastin, a new AAA protein, is altered in the most frequent form of autosomal dominant spastic paraplegia. Nature Genet. 23: 296-303, 1999. [PubMed: 10610178] [Full Text: https://doi.org/10.1038/15472]

  14. Hazan, J., Fontaine, B., Bruyn, R. P. M., Lamy, C., van Deutekom, J. C. T., Rime, C. S., Durr, A., Melki, J., Lyon-Caen, O., Agid, Y., Munnich, A., Padberg, G. W., de Recondo, J., Frants, R. R., Brice, A., Weissenbach, J. Linkage of a new locus for autosomal dominant familial spastic paraplegia to chromosome 2p. Hum. Molec. Genet. 3: 1569-1573, 1994. [PubMed: 7833913] [Full Text: https://doi.org/10.1093/hmg/3.9.1569]

  15. Hazan, J., Lamy, C., Melki, J., Munnich, A., de Recondo, J., Weissenbach, J. Autosomal dominant familial spastic paraplegia is genetically heterogeneous and one locus maps to chromosome 14q. Nature Genet. 5: 163-167, 1993. [PubMed: 8252041] [Full Text: https://doi.org/10.1038/ng1093-163]

  16. Hentati, A., Pericak-Vance, M. A., Lennon, F., Wasserman, B., Hentati, F., Juneja, T., Angrist, M. H., Hung, W.-Y., Boustany, R.-M., Bohlega, S., Iqbal, Z., Huether, C. H., Ben Hamida, M., Siddique, T. Linkage of a locus for autosomal dominant familial spastic paraplegia to chromosome 2p markers. Hum. Molec. Genet. 3: 1867-1871, 1994. [PubMed: 7849714] [Full Text: https://doi.org/10.1093/hmg/3.10.1867]

  17. Hsu, D. R., Meyer, B. J. The dpy-30 gene encodes an essential component of the Caenorhabditis elegans dosage compensation machinery. Genetics 137: 999-1018, 1994. [PubMed: 7982580] [Full Text: https://doi.org/10.1093/genetics/137.4.999]

  18. Ishiura, H., Takahashi, Y., Hayashi, T., Saito, K., Furuya, H., Watanabe, M., Murata, M., Suzuki, M., Sugiura, A., Sawai, S., Shibuya, K., Ueda, N., Ichikawa, Y., Kanazawa, I., Goto, J., Tsuji, S. Molecular epidemiology and clinical spectrum of hereditary spastic paraplegia in the Japanese population based on comprehensive mutational analyses. J. Hum. Genet. 59: 163-172, 2014. [PubMed: 24451228] [Full Text: https://doi.org/10.1038/jhg.2013.139]

  19. Lindsey, J. C., Lusher, M. E., McDermott, C. J., White, K. D., Reid, E., Rubinsztein, D. C., Bashir, R., Hazan, J., Shaw, P. J., Bushby, K. M. D. Mutation analysis of the spastin gene (SPG4) in patients with hereditary spastic paraparesis. J. Med. Genet. 37: 759-765, 2000. [PubMed: 11015453] [Full Text: https://doi.org/10.1136/jmg.37.10.759]

  20. McDermott, C. J., Burness, C. E., Kirby, J., Cox, L. E., Rao, D. G., Hewamadduma, C., Sharrack, B., Hadjivassiliou, M., Chinnery, P. F., Dalton, A., Shaw, P. J. Clinical features of hereditary spastic paraplegia due to spastin mutation. Neurology 67: 45-51, 2006. Note: Erratum: Neurology 72: 1534 only, 2009. [PubMed: 16832076] [Full Text: https://doi.org/10.1212/01.wnl.0000223315.62404.00]

  21. McMonagle, P., Byrne, P. C., Fitzgerald, B., Webb, S., Parfrey, N. A., Hutchinson, M. Phenotype of AD-HSP due to mutations in the SPAST gene: comparison with AD-HSP without mutations. Neurology 55: 1794-1800, 2000. [PubMed: 11134375] [Full Text: https://doi.org/10.1212/wnl.55.12.1794]

  22. McMonagle, P., Byrne, P., Hutchinson, M. Further evidence of dementia in SPG4-linked autosomal dominant hereditary spastic paraplegia. Neurology 62: 407-410, 2004. [PubMed: 14872021] [Full Text: https://doi.org/10.1212/01.wnl.0000108629.04434.05]

  23. Meijer, I. A., Hand, C. K., Cossette, P., Figlewicz, D. A., Rouleau, G. A. Spectrum of SPG4 mutations in a large collection of North American families with hereditary spastic paraplegia. Arch. Neurol. 59: 281-286, 2002. [PubMed: 11843700] [Full Text: https://doi.org/10.1001/archneur.59.2.281]

  24. Mitne-Neto, M., Kok, F., Beetz, C., Pessoa, A., Bueno, C., Graciani, Z., Martyn, M., Monteiro, C. B. M., Mitne, G., Hubert, P., Nygren, A. O. H., Valadares, M., Cerqueira, A. M. P., Starling, A., Deufel, T., Zatz, M. A multi-exonic SPG4 duplication underlies sex-dependent penetrance of hereditary spastic paraplegia in a large Brazilian pedigree. Europ. J. Hum. Genet. 15: 1276-1279, 2007. [PubMed: 17895902] [Full Text: https://doi.org/10.1038/sj.ejhg.5201924]

  25. Miura, S., Shibata, H., Kida, H., Noda, K., Toyama, T., Iwasaki, N., Iwaki, A., Ayabe, M., Aizawa, H., Taniwaki, T., Fukumaki, Y. Partial SPAST and DPY30 deletions in a Japanese spastic paraplegia type 4 family. Neurogenetics 12: 25-31, 2011. [PubMed: 20857310] [Full Text: https://doi.org/10.1007/s10048-010-0260-7]

  26. Murphy, S., Gorman, G., Beetz, C., Byrne, P., Dytko, M., McMonagle, P., Kinsella, K., Farrell, M., Hutchinson, M. Dementia in SPG4 hereditary spastic paraplegia: clinical, genetic, and neuropathologic evidence. Neurology 73: 378-384, 2009. [PubMed: 19652142] [Full Text: https://doi.org/10.1212/WNL.0b013e3181b04c6c]

  27. Nance, M. A., Raabe, W. A., Midani, H., Kolodny, E. H., David, W. S., Megna, L., Pericak-Vance, M. A., Haines, J. L. Clinical heterogeneity of familial spastic paraplegia linked to chromosome 2p21. Hum. Hered. 48: 169-178, 1998. [PubMed: 9618065] [Full Text: https://doi.org/10.1159/000022798]

  28. Nielsen, J. E., Koefoed, P., Abell, K., Hasholt, L., Eiberg, H., Fenger, K., Niebuhr, E., Sorensen, S. A. CAG repeat expansion in autosomal dominant pure spastic paraplegia linked to chromosome 2p21-p24. Hum. Molec. Genet. 6: 1811-1816, 1997. [PubMed: 9302257] [Full Text: https://doi.org/10.1093/hmg/6.11.1811]

  29. Nielsen, J. E., Krabbe, K., Jennum, P., Koefoed, P., Jensen, L. N., Fenger, K., Eiberg, H., Hasholt, L., Werdelin, L., Sorensen, S. A. Autosomal dominant pure spastic paraplegia: a clinical, paraclinical, and genetic study. J. Neurol. Neurosurg. Psychiat. 64: 61-66, 1998. [PubMed: 9436729] [Full Text: https://doi.org/10.1136/jnnp.64.1.61]

  30. Orlacchio, A., Gaudiello, F., Totaro, A., Floris, R., St George-Hyslop, P. H., Bernardi, G., Kawarai, T. A new SPG4 mutation in a variant form of spastic paraplegia with congenital arachnoid cysts. Neurology 62: 1875-1878, 2004. [PubMed: 15159500] [Full Text: https://doi.org/10.1212/01.wnl.0000125324.32082.d9]

  31. Orlacchio, A., Kawarai, T., Totaro, A., Errico, A., St George-Hyslop, P. H., Rugarli, E. I., Bernardi, G. Hereditary spastic paraplegia: clinical genetic study of 15 families. Arch. Neurol. 61: 849-855, 2004. [PubMed: 15210521] [Full Text: https://doi.org/10.1001/archneur.61.6.849]

  32. Orlacchio, A., Patrono, C., Gaudiello, F., Rocchi, C., Moschella, V., Floris, R., Bernardi, G., Kawarai, T. Silver syndrome variant of hereditary spastic paraplegia: a locus to 4p and allelism with SPG4. Neurology 70: 1959-1966, 2008. [PubMed: 18401025] [Full Text: https://doi.org/10.1212/01.wnl.0000294330.27058.61]

  33. Reid, E., Grayson, C., Rubinsztein, D. C., Rogers, M. T., Rubinsztein, J. S. Subclinical cognitive impairment in autosomal dominant 'pure' hereditary spastic paraplegia. J. Med. Genet. 36: 797-798, 1999. [PubMed: 10528866] [Full Text: https://doi.org/10.1136/jmg.36.10.797]

  34. Sack, G. H., Jr., Huether, C. A., Garg, N. Familial spastic paraplegia--clinical and pathologic studies in a large kindred. Johns Hopkins Med. J. 143: 117-121, 1978. [PubMed: 703033]

  35. Scott, W. K., Gaskell, P. C., Lennon, F., Wolpert, C. M., Menold, M. M., Aylsworth, A. S., Warner, C., Farrell, C. D., Boustany, R.-M. N., Albright, S. G., Boyd, E., Kingston, H. M., Cumming, W. J. K., Vance, J. M., Pericak-Vance, M. A. Locus heterogeneity, anticipation and reduction of the chromosome 2p minimal candidate region in autosomal dominant familial spastic paraplegia. Neurogenetics 1: 95-102, 1997. [PubMed: 10732810] [Full Text: https://doi.org/10.1007/s100480050014]

  36. Shoukier, M., Neesen, J., Sauter, S. M., Argyriou, L., Doerwald, N., Pantakani, D. V. K., Mannan, A. U. Expansion of mutation spectrum, determination of mutation cluster regions and predictive structural classification of SPAST mutations in hereditary spastic paraplegia. Europ. J. Hum. Genet. 17: 187-194, 2009. Note: Erratum: Europ. J. Hum. Genet. 17: 401-402, 2009. [PubMed: 18701882] [Full Text: https://doi.org/10.1038/ejhg.2008.147]

  37. Starling, A., Rocco, P., Passos-Bueno, M. R., Hazan, J., Marie, S. K., Zatz, M. Autosomal dominant (AD) pure spastic paraplegia (HSP) linked to locus SPG4 affects almost exclusively males in a large pedigree. J. Med. Genet. 39: e77, 2002. Note: Electronic article. [PubMed: 12471215] [Full Text: https://doi.org/10.1136/jmg.39.12.e77]

  38. Svenson, I. K., Ashley-Koch, A. E., Gaskell, P. C., Riney, T. J., Cumming, W. J. K., Kingston, H. M., Hogan, E. L., Boustany, R.-M. N., Vance, J. M., Nance, M. A., Pericak-Vance, M. A., Marchuk, D. A. Identification and expression analysis of spastin gene mutations in hereditary spastic paraplegia. Am. J. Hum. Genet. 68: 1077-1085, 2001. [PubMed: 11309678] [Full Text: https://doi.org/10.1086/320111]

  39. Svenson, I. K., Kloos, M. T., Gaskell, P. C., Nance, M. A., Garbern, J. Y., Hisanaga, S., Pericak-Vance, M. A., Ashley-Koch, A. E., Marchuk, D. A. Intragenic modifiers of hereditary spastic paraplegia due to spastin gene mutations. Neurogenetics 5: 157-164, 2004. [PubMed: 15248095] [Full Text: https://doi.org/10.1007/s10048-004-0186-z]

  40. White, K. D., Ince, P. G., Lusher, M., Lindsey, J., Cookson, M., Bashir, R., Shaw, P. J., Bushby, K. M. D. Clinical and pathologic findings in hereditary spastic paraparesis with spastin mutation. Neurology 55: 89-94, 2000. [PubMed: 10891911] [Full Text: https://doi.org/10.1212/wnl.55.1.89]


Contributors:
Cassandra L. Kniffin - updated : 4/21/2014
Cassandra L. Kniffin - updated : 2/12/2013
Cassandra L. Kniffin - updated : 12/17/2009
Cassandra L. Kniffin - updated : 8/28/2009
Cassandra L. Kniffin - updated : 10/1/2008
Cassandra L. Kniffin - updated : 3/19/2008
Cassandra L. Kniffin - updated : 8/20/2007
Cassandra L. Kniffin - updated : 7/24/2007
Cassandra L. Kniffin - updated : 4/27/2007
Cassandra L. Kniffin - updated : 2/6/2007
Cassandra L. Kniffin - updated : 4/11/2006
Cassandra L. Kniffin - updated : 1/26/2005
Cassandra L. Kniffin - updated : 10/26/2004
Cassandra L. Kniffin - updated : 8/30/2004
Cassandra L. Kniffin - updated : 7/27/2004
Cassandra L. Kniffin - updated : 12/27/2002
Cassandra L. Kniffin - reorganized : 10/4/2002
Cassandra L. Kniffin - updated : 6/26/2002
Victor A. McKusick - updated : 6/15/2001
George E. Tiller - updated : 4/14/2000
Michael J. Wright - updated : 1/19/2000
Victor A. McKusick - updated : 11/2/1998
Victor A. McKusick - updated : 10/2/1998
Victor A. McKusick - updated : 8/17/1998
Victor A. McKusick - updated : 7/7/1998
Victor A. McKusick - updated : 5/5/1998
Victor A. McKusick - updated : 11/4/1997
Victor A. McKusick - updated : 5/30/1997

Creation Date:
Victor A. McKusick : 10/14/1993

Edit History:
alopez : 11/17/2023
alopez : 11/17/2023
carol : 04/10/2023
carol : 10/09/2019
carol : 10/08/2019
carol : 08/31/2016
carol : 04/23/2014
ckniffin : 4/21/2014
tpirozzi : 8/14/2013
tpirozzi : 8/14/2013
tpirozzi : 8/14/2013
tpirozzi : 8/13/2013
terry : 3/15/2013
carol : 3/11/2013
carol : 3/11/2013
alopez : 2/18/2013
ckniffin : 2/12/2013
wwang : 3/2/2011
wwang : 1/8/2010
ckniffin : 12/17/2009
wwang : 9/14/2009
ckniffin : 8/28/2009
wwang : 11/11/2008
wwang : 10/3/2008
ckniffin : 10/1/2008
wwang : 4/10/2008
ckniffin : 3/19/2008
wwang : 9/6/2007
ckniffin : 8/20/2007
wwang : 8/3/2007
ckniffin : 7/24/2007
wwang : 6/8/2007
ckniffin : 4/27/2007
wwang : 2/8/2007
ckniffin : 2/6/2007
wwang : 4/20/2006
ckniffin : 4/11/2006
terry : 6/24/2005
tkritzer : 2/2/2005
ckniffin : 1/26/2005
tkritzer : 11/1/2004
ckniffin : 10/26/2004
carol : 9/7/2004
ckniffin : 8/30/2004
tkritzer : 7/28/2004
ckniffin : 7/27/2004
carol : 1/6/2003
ckniffin : 12/27/2002
carol : 10/4/2002
ckniffin : 9/30/2002
tkritzer : 8/9/2002
ckniffin : 6/26/2002
cwells : 6/27/2001
terry : 6/15/2001
alopez : 4/17/2000
terry : 4/14/2000
alopez : 1/19/2000
alopez : 11/9/1999
alopez : 11/2/1999
mgross : 9/24/1999
carol : 11/11/1998
terry : 11/2/1998
dkim : 10/12/1998
carol : 10/7/1998
terry : 10/2/1998
carol : 8/18/1998
terry : 8/17/1998
carol : 7/9/1998
terry : 7/7/1998
carol : 5/12/1998
terry : 5/5/1998
jenny : 11/12/1997
terry : 11/4/1997
alopez : 7/29/1997
terry : 7/7/1997
jenny : 6/3/1997
terry : 5/30/1997
mimadm : 3/25/1995
carol : 1/23/1995
terry : 11/16/1994
carol : 10/14/1993