Next Article in Journal
Preclinical Detection of Early Glomerular Injury in Children with Kidney Diseases—Independently of Usual Markers of Kidney Impairment and Inflammation
Previous Article in Journal
GDSL in Lilium pumilum (LpGDSL) Confers Saline–Alkali Resistance to the Plant by Enhancing the Lignin Content and Balancing the ROS
Previous Article in Special Issue
Navigating Lipodystrophy: Insights from Laminopathies and Beyond
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Lipodystrophic Laminopathies: From Dunnigan Disease to Progeroid Syndromes

by
Everardo Josué Díaz-López
1,2,
Sofía Sánchez-Iglesias
1,
Ana I. Castro
2,3,
Silvia Cobelo-Gómez
1,
Teresa Prado-Moraña
1,2,
David Araújo-Vilar
1,2 and
Antia Fernandez-Pombo
1,2,*
1
UETeM-Molecular Pathology Group, Department of Psychiatry, Radiology, Public Health, Nursing and Medicine, IDIS-CIMUS, University of Santiago de Compostela, 15706 Santiago de Compostela, Spain
2
Division of Endocrinology and Nutrition, University Clinical Hospital of Santiago de Compostela, 15706 Santiago de Compostela, Spain
3
CIBER Fisiopatología de la Obesidad y la Nutrición (CIBERobn), 28029 Madrid, Spain
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(17), 9324; https://doi.org/10.3390/ijms25179324
Submission received: 14 July 2024 / Revised: 20 August 2024 / Accepted: 26 August 2024 / Published: 28 August 2024
(This article belongs to the Special Issue Adipose Tissue Dynamics in Laminopathies)

Abstract

:
Lipodystrophic laminopathies are a group of ultra-rare disorders characterised by the presence of pathogenic variants in the same gene (LMNA) and other related genes, along with an impaired adipose tissue pattern and other features that are specific of each of these disorders. The most fascinating traits include their complex genotype-phenotype associations and clinical heterogeneity, ranging from Dunnigan disease, in which the most relevant feature is precisely adipose tissue dysfunction and lipodystrophy, to the other laminopathies affecting adipose tissue, which are also characterised by the presence of signs of premature ageing (Hutchinson Gilford-progeria syndrome, LMNA-atypical progeroid syndrome, mandibuloacral dysplasia types A and B, Nestor-Guillermo progeria syndrome, LMNA-associated cardiocutaneous progeria). This raises several questions when it comes to understanding how variants in the same gene can lead to similar adipose tissue disturbances and, at the same time, to such heterogeneous phenotypes and variable degrees of metabolic abnormalities. The present review aims to gather the molecular basis of adipose tissue impairment in lipodystrophic laminopathies, their main clinical aspects and recent therapeutic strategies. In addition, it also summarises the key aspects for their differential diagnosis.

1. Introduction

Nuclear lamins are crucial nuclear envelope proteins that provide essential structural support and facilitate interactions between extranuclear structures and nucleoplasm components. The A-type nuclear lamins A and C are encoded by the LMNA gene [1,2].
Laminopathies are inherited diseases resulting from abnormalities in A-type lamins due to variants in the LMNA gene or in other genes involved in prelamin A processing [3,4]. These disorders have diverse clinical manifestations, which may include bone and cardiac abnormalities, lipodystrophy, dermopathy, neuropathy, metabolic abnormalities and premature ageing [5]. In particular, lipodystrophic laminopathies, characterised by selective loss of adipose tissue, represent a distinct subset of these conditions [6].
There are several diseases related to the LMNA gene presenting with adipose tissue dysfunction, such as Dunnigan disease or familial partial lipodystrophy type 2 (FPLD2), Hutchinson-Gilford progeria syndrome (HGPS), LMNA-associated atypical progeroid syndrome (APS) and mandibuloacral dysplasia type A (MADA) [7]. Among the progeroid syndromes, those associated with LMNA variants are particularly significant due to their resemblance to aspects of ageing and range from the mild acceleration observed in MADA to severe acceleration in HGPS [8]. In addition, variants in genes affecting lamin A/C processing and nuclear lamina structure, such as ZMPSTE24 (linked to mandibuloacral dysplasia type B [MADB]) and BANF1 (associated with Nestor-Guillermo progeria syndrome [NGPS]), can also result in accelerated ageing syndromes that share adipose tissue impairment [9].
These syndromes underscore the critical role of nuclear envelope proteins in human health and require an interdisciplinary clinical approach to improve diagnosis, monitoring and therapeutic development. This review aims to comprehensively examine lipodystrophic laminopathies, covering the genetic basis of adipose tissue disturbance as a common denominator of these disorders, their clinical manifestations and current therapeutic strategies, in order to improve understanding and guide future research and clinical practice.

2. Aetiopathogenesis of Adipose Tissue Impairment in Lipodystrophic Laminopathies

The nuclear lamina is a dense fibrillar network located underneath the inner nuclear membrane. Major components of the nuclear lamina are members of the lamin family of type V intermediate filament proteins, which are important determinants of nuclear and cellular architecture and are significant regulators of stem cell differentiation [10]. The LMNA gene codifies for A-type lamins and LMNB1 and LMNB2 for lamins B1 and B2, respectively. In this sense, prelamin A is farnesylated, methylated and processed by ZMPSTE24 to form the mature lamina. The correct processing of the A-type lamins is essential for the prevention of laminopathies, which are, as previously mentioned, caused by variants in the LMNA gene (such as Dunnigan disease or several premature ageing syndromes), in other genes that influence lamina processing (such as ZMPSTE24 gene variants in MADB or in restrictive dermopathy) or in genes that influence its proper functioning on chromatin (such as BANF1 gene variants in NGPS). Thus, LMNA-related syndromes are considered to be primary laminopathies and the others secondary laminopathies [10,11,12].
Laminopathies mainly affect mesenchymal tissues (adipose tissue, muscle, bone) [10], although, due to their complex genotype-phenotype associations, not all of them share the same mesenchymal tissue abnormalities. Thus, adipose tissue is affected specifically in the primary lipodystrophic laminopathies Dunnigan disease, HGPS, LMNA-atypical progeroid syndrome (APS) and MADA, and in the secondary lipodystrophic laminopathies MADB and NGPS [11]. This raises several questions when it comes to understanding how variants in the same gene can lead to similar adipose tissue disturbances and, at the same time, to such heterogeneous phenotypes.
Thus, the precise mechanisms linking nuclear envelope abnormalities to lipodystrophy remain unknown. However, various theories have been discussed over the years and different pathogenetic mechanisms leading to improper fat distribution in lamin A-linked lipodystrophies have been proposed [10,13,14,15,16,17,18,19,20,21,22,23,24,25]. One such theory is that regulated structural reorganisation of A-type lamins is needed during mesenchymal stem cell differentiation or quiescence. If these rearrangements in the organisation of the lamina are hindered by mutant lamins or their abscence, they could ultimately lead to adipose tissue or muscle impairment [20]. Another theory refers to gene expression and an alteration in transcription factors such as the sterol regulatory element-binding protein 1c (SREBP-1c), which, when sequestered by prelamin A, is not able to properly activate the peroxisome proliferator-activated receptor gamma (PPARG) [18], which is essential in the adipogenesis process. On the other hand, variants in the LMNA gene may prevent the correct maturation of prelamin A, generating aberrant electrical charges that alter the interaction with several peptides, such as SREBP-1c itself [26], precluding adipogenesis through the previously mentioned reduction of PPARG activation [27]. Other factors could intervene in this inactivation of PPARG, such as the barrier-to-autointegration factor (BAF) [28] or the retinoblastoma protein (pRb) [29], whose subnuclear localisation and degradation by the proteasome also depend on A-type lamins. pRB promotes adipogenesis, and its phosphorylated form is reduced in lipoatrophic areas in patients with Dunnigan disease [30]. Furthermore, the accumulation of prelamin A alters the nuclear envelope and chromatin organisation, generating oxidative stress, the accumulation of reactive oxygen species [30,31] and a premature ageing phenotype [14,30]. As for the reason why mutant lamin A is unable to complete its maturation, the altered electrical charges previously mentioned were considered responsible by preventing the correct coupling of ZMPSTE24 with prelamin A [32].
Another mechanism proposed to explain the alteration in adipogenesis and premature ageing is the alteration in autophagic flux, which seemed to be reduced [33,34]. Early activation of autophagy in laminopathic adipocyte precursors followed by autophagic flux impairment has also been related to impaired white adipocyte turnover and failure of adipose tissue browning and, therefore, to fat loss and improper accumulation in these syndromes [23]. However, the reason why some areas of the body can be affected and not others, with lamins being ubiquitous proteins, is not completely clear. This could be explained due to the differential expression of different variants of lamins, originated by alternative splicing, establishing different relationships and exerting different functions, depending on the cell type [35].
Nevertheless, the clinical heterogeneity of these disorders also suggests the possible influence of other factors that may even induce epigenetic changes. Thus, one of the hypotheses being raised is that variability in the clinical expressivity of these disorders could be triggered by certain epigenetic marks. Moreover, some studies suggest that the mechanisms responsible for the loss of adipose tissue may be related to certain microRNAs (miRNAs) and histone modifications. In this sense, miRNAs are small, non-coding RNA molecules that play a crucial role in the post-transcriptional regulation of gene expression, usually by mRNA cleavage, mRNA destabilisation or translational repression. They have been shown to be involved in important epigenetic mechanisms such as histone modifications, which are relevant for chromatin formation and the maintenance of the nuclear lamina [21,36,37]. Specifically, the role of dynamic chromatin remodelling on adipogenesis has recently been reported and, more precisely, how pathogenic variants in the LMNA gene can modulate certain epigenetic marks such as the anti-adipogenic miRNA miR-335 in adipocyte progenitors, altering the processes of adipocyte differentiation [21].

3. Dunnigan Disease

Dunnigan disease, also known as FPLD2 (MIM#151660), is characterised by the loss of adipose tissue from the trunk, buttocks and upper and lower limbs mainly from puberty, associated with an accumulation of fat in the face, neck, supraclavicular fossae and labia majora/pubic area [38,39,40,41]. Most of the subjects with this phenotype harbour heterozygous missense variants affecting arginine at codon 482 in exon 8 of the LMNA gene. However, variants in other exons have also been reported and are shown to promote atypical FPLD2, with a variable pattern and severity of fat loss [42,43,44,45,46,47,48,49,50,51] (Figure 1 and Figure 2). There is also a tendency towards the formation of subcutaneous lipomas (found in 20% of cases), which can guide the diagnosis in subjects with a concordant phenotype [30] and, along with this reported abnormal pattern of fat distribution, muscular hypertrophy and myalgias are also recurrent features [52,53,54]. Identification of the “Dunnigan sign” (hypertrophy of mons pubis fat surrounded by subcutaneous lipoatrophy) is likewise helpful in diagnosing women with this disorder. Due to their lack of subcutaneous fat, these patients tend to show very prominent peripheral veins [55]. Signs of hyperandrogenism, such as hirsutism, can be seen in affected women [52,56], as well as the presence of signs of insulin resistance, including acrochordons and acanthosis nigricans [57,58,59] (Figure 1). The clinical diagnosis of this disorder is not difficult in women for trained physicians considering that the expression and severity of the previously mentioned phenotype may be markedly dependent on sex, with women being more affected. On the contrary, men typically show a less-apparent lipodystrophy phenotype and a later onset [41,42].
Dunnigan disease can become a serious disorder with potentially lethal associated comorbidities and metabolic complications. Although these comorbidities usually develop after the age of 10, an anticipation phenomenon has been described, with the occurrence of metabolic disturbances at an earlier age across generations [60,61,62]. In this sense, the response of maladaptive adipose tissue acquires a dysfunctional pattern of adipocytokine production. These adipocytokines, synthesised both by adipocytes and by other inflammatory cells, interfere with insulin signalling pathways, generating resistance to its action [63]. Ectopic fat accumulation, mainly in the liver and muscle, also plays a relevant role in the aetiopathogenesis of insulin resistance. This can lead to a variable degree of metabolic complications, including non-ketotic diabetes, with a prevalence ranging from 28% to 51% [58,64,65,66,67,68], or dyslipidaemia, which is fundamentally characterised by high triglyceride levels, also related to acute pancreatitis in some of these subjects [69], and low high-density lipoprotein cholesterol [46,58,61,70,71,72,73]. Hypertension has also been described with a prevalence of up to 41% [46,59,74]. Kidney alterations including proteinuria and renal failure have likewise been reported [75]. Non-alcoholic fatty liver disease and non-alcoholic steatohepatitis, ultimately leading to cirrhosis, are other frequent organ abnormalities due to ectopic lipid storage in the liver [58,76]. An increased prevalence of cardiac manifestations, such as cardiac hypertrophy [56,77], atrioventricular conduction defects, heart failure due to ventricular dilatation [78,79], early atherosclerosis [80,81,82] and arrhythmias [78,79,83], has also been described in this population. Polycystic ovary syndrome, fertility problems and a higher rate of miscarriages and stillbirths can likewise be observed [84,85,86,87].
Furthermore, it should be taken into account that Dunnigan disease, in addition to belonging to the laminopathy family due to the involvement of the LMNA gene in its aetiopathogenesis, also belongs to the FPLD group of syndromes, which comprises a total of nine FPLD subtypes and another three unclassified variants of FPLD [88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110]. This set of disorders shares most of the previously mentioned phenotypic characteristics and comorbidities and, therefore, attention should be paid to certain peculiarities of each specific disease that may help to perform a correct differential diagnosis. The main differential characteristics of Dunnigan disease and the other 11 FPLD subtypes can be seen in Table 1.
There is currently no cure for this disorder and fat loss is generally not recovered. However, the morbidity and mortality of this condition could be improved with early intervention and, therefore, its treatment should be fundamentally oriented towards the control of the metabolic complications previously described. Apart from diet, along with physical activity (after an adequate cardiological assessment and avoiding vigorous exercise) and standard treatments mainly aimed at the control of diabetes and dyslipidaemia, several drugs have been specifically tested in this population. Metreleptin, a human recombinant leptin, has been approved by the European Medicines Agency (EMA) for the treatment of FPLD subjects > 12 years of age who have not responded to standard therapies (glycated haemoglobin [HbA1c] > 8% and/or triglycerides > 500 mg/dL) (https://www.ema.europa.eu/en/medicines/human/EPAR/myalepta, accessed on 7 July 2024). However, its response in partial lipodystrophy is considered to be less robust than in generalised forms. In this sense, while metreleptin has been proven to reduce serum triglyceride levels regardless of patients’ leptin levels, the reduction in HbA1c was found to be inconsistent [111,112,113]. It also seems to improve hepatic steatosis [114], although more studies in this field and in this specific population are needed. On the other hand, new drugs focused on hepatic lipid metabolism have also been investigated in recent years with promising results, as is the case of volanesorsen, an antisense inhibitor of apolipoprotein C-III. Its efficacy and safety have been evaluated in a 52-week phase II/III study, showing an 88% decrease in triglycerides and a significant reduction in hepatic fat fraction in 40 patients with FPLD [115]. Less robust results have been observed for gemcabene calcium, a monocalcium salt of a dialkyl ether dicarboxylic acid, the efficacy and safety of which has been evaluated in five women with FPLD (two with the LMNA R482Q variant), showing a median change in serum triglycerides of −19.6% [116]. Vupanorsen, an inhibitor of ANGPTL3, has also been studied in a small number of patients with FPLD (two of them with Dunnigan disease) showing a reduction of triglyceride fasting levels of 59.9%, as well as in other lipoproteins [117]. In addition, a phase II study regarding the effectiveness of obeticholic acid in reducing hepatic triglyceride levels in patients with Dunnigan disease and hepatic steatosis has recently been completed (ClinicalTrials. gov Identifier: NCT02430077).

4. Hutchinson-Gilford Progeria Syndrome

With an estimated prevalence of one case per 20 million inhabitants (www.progeriaresearch.org; accessed on 7 July 2024), HGPS (#MIM176670) is considered to be one of the most severe laminopathies, originating from the aberrant splicing of the LMNA gene and the consequent expression of an abnormal lamin A protein, called progerin [118,119]. Patients are healthy at birth with the phenotype usually becoming evident at 18–24 months of age, including a broad range of clinical features [120,121,122] (Figure 1 and Figure 2): low body weight, with severe generalised lipodystrophy while preserving intraabdominal fat, along with muscular atrophy, growth retardation and short stature, early alopecia with prominent scalp veins, loss of skin elasticity and skin hypo/hyperpigmentation, beaked nose, micrognathia, high-arched palate, mandibular osteolysis, dental crowding, high-pitched voice, sensorineural hearing loss, nail dystrophy, joint stiffness, short club-shaped distal phalanges (acroosteolysis) and osteopenia. Patients with HGPS do not show intellectual impairment [119].
The most significant abnormalities in HGPS, which ultimately lead to death at around the age of 14 [123], are cardiovascular complications (myocardial ischemia, infarction and stroke) [124]. Patients can suffer insulin resistance with some showing elevated levels of serum triglycerides, total cholesterol and low-density lipoprotein cholesterol with reduced levels of high-density lipoprotein cholesterol [122,125]. However, interestingly, metabolic derangements are mild and HGPS patients rarely develop dyslipidaemia or increased C-reactive protein, two characteristics often observed in cardiovascular disease in the general population [124,126]. This leads to the suspicion that the mechanisms leading to the development of these cardiovascular complications in this group of patients differ from the normal ageing population. Thus, beyond atherosclerosis, there is evidence of intimal thickening and an increase in arterial stiffness leading to many measurable changes in the vasculature of these patients [122,127]. They also present vascular walls with a dramatically thickened fibrotic matrix with a depletion of medial smooth muscle cells that is replaced by collagen, with secondary maladaptive vascular remodelling, probably due in part to the extreme sensitivity of these cells to progerin expression [122,127,128,129,130].
Even though there is currently no cure for this disorder, in recent years, a number of therapies have shown promise in preclinical stages for its treatment and, in 2020, lonafarnib became the first (and only) ever U.S. Food and Drug Administration (FDA) and EMA approved treatment for this condition. This drug is a farnesyltransferase inhibitor, which works by inhibiting the processing of prelamin A to progerin [131,132,133], the administration of which has been shown to improve some of the symptoms of the disease (such as rate of weight gain, pulse-wave velocity, carotid artery wall echodensity, skeletal rigidity, cardiovascular stiffness, bone density and sensorineural hearing) and also a decreased mortality rate [132,134,135]. In combination with lonafarnib, the rapamycin analog everolimus is also currently being tested in an ongoing phase I/II clinical trial (ClinicalTrials.gov Identifier: NCT02579044). Likewise, particularly remarkable is the addition of progerinin as a new candidate specifically developed for progeria, the safety, tolerability, pharmacokinetics and pharmacodynamic profile of which is being studied in another phase I clinical trial (ClinicalTrials.gov Identifier: NCT04512963). Furthermore, a potential genetic therapeutic strategy employing antisense peptide-conjugated phosphorodiamidate morpholino oligomers (PPMOs) to block the pathogenic splicing of mutant transcripts has shown a significant reduction of progerin transcripts in the aorta, a 61.6% increase in lifespan and rescue of vascular smooth muscle cell loss in large arteries in transgenic mouse models of HGPS [136]. CRISPR/Cas9 gene editing also seems a promising strategy for the treatment of genetic diseases, including HGPS [137].

5. LMNA-Atypical Progeroid Syndrome

This syndrome, due to different missense LMNA variants (such as P4R, E111K, D136H, E159K, C588R, R349W or T10I) is characterised by a marked phenotypic heterogeneity, with evidence of both partial and generalised lipodystrophy (Figure 2). While the accumulation of farnesylated prelamin A has been demonstrated in some cases, in other cases the pathogenesis of clinical manifestations was not related to this accumulation [138,139,140].
The onset of the phenotype occurs later than in HGPS and MAD, during childhood or early adulthood, [139,141,142] and life expectancy is likewise longer [143]. Patients with APS usually share several progeroid features (Figure 1), including short stature, partial alopecia, early greying of hair, mottled skin pigmentation, sclerodermiform lesions, beaked nose, high-arched palate, micrognatia, abnormal teeth implantation with dental crowding, high-pitched voice, sensorineural hearing impairment, joint stiffness and osteoporosis [138,139,144]. However, although they have overlapping features when compared with HGPS and MAD patients, acroosteolysis and clavicular resorption/hypoplasia are usually absent or mild in this specific disorder [139]. On the other hand, this syndrome is frequently associated with metabolic abnormalities and cardiovascular complications, including valvular disease, rhythm disturbances, coronary artery disease and cardiomyopathy. Proteinuric nephropathy is also present in the majority of cases [139,142,144,145].
In 2018, Hussain I et al. proposed the designation of a distinctive syndrome due to heterozygous LMNA p.T10I variants, called generalised lipodystrophy-associated progeroid syndrome, considering the unique and relatively homogeneous clinical features of this disorder in comparison with the previously reported APS. These specific characteristics were early childhood onset of generalised lipodystrophy along with other progeroid features, more severe metabolic complications and a notable need for cardiac transplantation [140]. Another recurrent variant is the previously mentioned R349W, which leads to a multisystem progeroid syndrome with lipodystrophy characterised by a loss of fat in the limbs and face and its accumulation in the dorsocervical region, along with the shared progeroid features described [141,142]. On the other hand, the missense LMNA variant p.(Thr528Met) was identified in heterozygosity in subjects with FPLD2, in a compound heterozygous state in subjects with APS and severe partial lipodystrophy and, recently, in homozygosity in subjects with homogeneous APS clinical features with major musculoskeletal involvement [146].
Regarding the availability of specific therapies, it should be noted that metreleptin has been successfully used in some isolated cases for the treatment of associated comorbidities [147].

6. Other LMNA-Associated Lipodystrophies

6.1. LMNA-Associated Generalised Lipodystrophy

Only one case with the p.(Arg582His) variant in exon 11 of the LMNA gene in a homozygous fashion and near-total fat loss has been reported. However, adipose tissue was found to be preserved in the retroorbital area, mons pubis and the genital region. This homozygous variant was also associated with an earlier onset of metabolic complications in comparison to heterozygous cases [148]. Another family of four with the p.(Arg582Cys) variant in exon 11 of the LMNA gene in the homozygous and heterozygous states has been described. All members showed extremely distinct features including a severe generalised lipodystrophic phenotype and metabolic abnormalities leading to death at a young age [47]. Finally, although the pathogenic p.(Arg571Ser) variant in the LMNA gene has been previously associated with cardiomyopathy and neuropathy without lipodystrophy [149,150], in 2017, another two cases harbouring generalised loss of fat were also described. The onset of the phenotype in these cases occurred after birth and in early childhood with both subjects developing severe metabolic abnormalities and dying at an early age [151].

6.2. LMNA-Associated Cardiocutaneous Progeria

Only one patient has been reported with this specific syndrome due to the p.(Asp300Gly) variant in the LMNA gene developing an unspecified pattern of fat loss around the third decade of life. Cardiac and cutaneous manifestations along with a prematurely aged appearance were also described for this patient. In addition, it should be noted that this disorder may also be related to a greater incidence of malignancy, an unprecedented finding in LMNA-linked progeria disorders [152].

7. Mandibuloacral Dysplasia

Patients with MAD present with a range of clinical manifestations. Facial features typically include prominent scalp veins, ocular proptosis, beak-like nose, dental crowding and progressive osteolysis affecting the mandible, terminal phalanges and clavicles. Skeletal abnormalities include acroosteolysis, prominent interphalangeal joints and dystrophic nails [153]. Cutaneous manifestations in MAD include thin, wrinkled skin with mottled hyperpigmentation and atrophy in the acral regions, along with visible veins and tendons due to the absence of subcutaneous fat [154]. Although skeletal and cardiac muscles are usually unaffected in most cases of MAD, muscle weakness may occur with specific gene variants [155]. Features of metabolic syndrome such as insulin resistance, glucose intolerance, and hypertriglyceridemia have also been documented among affected individuals [156].
Two distinct patterns of body fat distribution are observed in patients with MAD-associated lipodystrophy. The first pattern (type A) involves loss of fat from the limbs along with normal or increased accumulation of fat in the face, neck and trunk. In contrast, the second pattern (type B) is characterised by a generalised loss of subcutaneous fat. A notable phenotypic difference lies in the specific association of LMNA variants with partial lipodystrophy (MADA) and ZMPSTE24 variants with generalised lipodystrophy (MADB) [157,158].

7.1. Mandibuloacral Dysplasia Type A

MADA (#MIM 248370) was first described in 1971 [159] as a rare autosomal recessive disorder resulting from abnormalities in nuclear lamin proteins, specifically lamins A and C [160]. MADA is predominantly caused by homozygous variants in the LMNA gene, leading to the accumulation of mutated prelamin A and lamin A/C [161]. Although most cases involve homozygous LMNA variants, there are instances of less frequent heterozygous compound conditions [162].
Symptoms typically appear in early childhood, becoming more pronounced during adolescence [153]. The clinical presentation of MADA exhibits significant homogeneity, despite its diverse impact on multiple organ systems among individuals [161]. A proposed classification categorises major clinical features seen in over 75% of patients, including acroosteolysis (100%), lipodystrophy (98%), mandibular hypoplasia (95%), clavicular hypoplasia (93%), growth retardation (79%) and a beaked nose (77%). Manifestations observed in 50–75% of patients include mottled skin pigmentation (72%), prominent cheeks (70%), prominent eyes (65%) and dental crowding (63%). Alopecia is a less common sign, reported in approximately half of the patients [163]. Female MADA patients may also present with lack of breast development and irregular menstrual periods [164].
MADA patients commonly exhibit lipodystrophy, characterised by partial loss of subcutaneous fat in the limbs and excessive fat deposition in the neck and trunk. In addition, variability in neck fat distribution is noted among patients, with some exhibiting normal fat levels while others develop fat accumulation, resembling a buffalo hump [14]. This pattern of fat distribution is reminiscent of that observed in FPLD2, also caused by LMNA gene variants [14,158]. In addition to skeletal and adipose tissue abnormalities, MADA is associated with mildly accelerated ageing, generalised joint stiffness and bone abnormalities that progress with age [165,166].
Overall, the comprehensive understanding of MADA requires ongoing research and clinical vigilance to manage its varied manifestations effectively. Current treatment focuses on managing the symptoms and complications associated with the condition, as there are no specifically approved therapies. Management should involve a multidisciplinary approach to address various aspects of the disorder, such as skeletal abnormalities, skin changes and potential cardiac issues.
Various cellular treatments have been developed to enhance the phenotype of cultured MADA cells by reducing accumulated prelamin A. Statins and farnesyl transferase inhibitors (FTIs) inhibit farnesylated prelamin A accumulation. Treatment of MADA cultured cells with statins was effective in the recovery of the chromatin phenotype, which is altered in MADA cells, but only in low passage cultures [161]. Combined treatment with mevinolin and trichostatin A improved the effects observed with mevinolin alone [161]. In addition, MADA fibroblasts subjected to chloroquine treatment showed increased levels of prelamin A, while impairing the autophagic process, suggesting an autophagic mechanism for removal of mutated prelamin A in these cells [8]. Moreover, inhibition of the mTOR pathway by rapamycin triggers lysosomal degradation of farnesylated prelamin A in MADA fibroblasts and rescues markers of cellular senescence as well as chromatin epigenetic mechanisms [167].
On the other hand, as previously described, patients affected by MADA suffer from an osteolytic process due to an excessive amount of released TGF-β2. Neutralising antibodies against TGF-β2, as well as statins or rapamycin, have been shown to block this osteolytic process, indicating potential therapeutic tools for MADA treatment. Everolimus treatment significantly improves the mutant phenotype and its pathogenetic pathways, suggesting it could be a therapeutic approach for MADA [168]. In this context, cytokines and other circulating molecules warrant further investigation as potential therapeutic targets for this syndrome [153].

7.2. Mandibuloacral Dysplasia Type B

MADB (#MIM 608612) is a rare premature ageing disorder following an autosomal recessive inheritance pattern [157]. MADB is primarily caused by variants in the ZMPSTE24 gene, resulting in impaired enzymatic activity and accumulation of farnesylated prelamin A [169].
Variants in the ZMPSTE24 gene have been associated with progeroid syndromes, including MADB and restrictive dermopathy, as well as a severe metabolic syndrome with abnormal fat accumulation and dilated cardiomyopathy [170]. However, patients with MADB typically present a more generalised loss of subcutaneous fat [171]. The median age at diagnosis is significantly younger (approximately 4 months) than subjects with MADA (approximately 4 years), suggesting a more severe phenotype [157].
The proposed diagnostic criteria for MADB require the presence of four or more major clinical criteria or the presence of three or more major clinical criteria and two or more minor clinical criteria. Major criteria include: short stature (height less than 2 standard deviations), clavicular hypoplasia, delayed closure of cranial sutures, high palate and/or mandibular hypoplasia and/or dental crowding, acroosteolysis of the distal phalanges (hands and/or feet), hypoplastic nails and/or brittle or sparse hair, and at least two of the following skin abnormalities: mottled pigmentation, atrophic skin, sclerodermic skin, or calcified skin nodules. Minor criteria include: lipoatrophy (generalised or partial) of the limbs, joint contractures and shortened phalanges. Major criteria are present in 85–100% of patients reported with MADB, while minor criteria are present in 70–84% of patients reported with MADB [172]. Failure of ossification of the interparietal region of the occipital bone is also a possible pathognomonic sign for MADB [173].
The differential diagnosis between MADA and MADB may be difficult especially in childhood, taking into account that the main clinical features are common [174]. A notable difference between MADA and MADB is the presence of renal disease in the latter group, such as focal segmental glomerulosclerosis and microhematuria [157]. Acanthosis nigricans and other metabolic disturbances have not been reported in MADB [174].
The early diagnosis and slowly progressive nature of MADB offers an opportunity to prevent the complications of the disease, such as renal disease. However, no specific therapeutic options are available. In mice, FTIs and combinations of statins with bisphosphonates have been tested to inhibit farnesylated prelamin A accumulation [131]. Treatment with zoledronate and pravastatin has improved various health indicators and increased median survival from 101 to 179 days [175]. Intravenous pamidronate infusion in a patient with MADB increased bone density but did not prevent osteolysis effectively [160]. Baricitinib alone or combined with FTIs showed beneficial effects on adipogenesis, suggesting potential therapeutic benefits pending further in vivo validation [176]. Lonafarnib corrected nuclear morphology in MADB patient cells, suggesting potential therapeutic benefits despite contradictory findings in some studies [177,178]. Treatment with rapamycin also showed promise in fibroblasts from ZMPSTE24-variant carriers, suggesting potential therapeutic avenues [177]. Morpholino antisense oligonucleotides targeting prelamin A reduced senescence markers and improved nuclear abnormalities in MADB patient fibroblasts [179]. Future strategies for premature ageing diseases such as MADB may include gene therapy (CRISPR), cell-based therapies focused on vascular tissue or RNA interference-based therapy [179]. Overall, the optimal therapeutic approach for MADB requires further investigation.

8. Nestor-Guillermo Progeria Syndrome

NGPS (#MIM 614008) is a rare autosomal recessive disorder. Genetic studies have identified the causative gene as Barrier-to-Autointegration Factor 1 (BANF1), although the reliability of the data remains limited due to the small number of reported cases (only three) [12,180,181]. Recently, another case in heterozygosis has also been reported, showing neuropathy and low weight [182].
Affected individuals appear normal at birth and reach normal developmental milestones. However, by the age of two years, symptoms such as growth retardation and lipodystrophy gradually begin to manifest, although in one case they presented at the age of one year [180]. The syndrome progresses with a marked decrease in mobility affecting both small and large joints over the following years. Hair loss begins during childhood and severe osteolysis becomes evident, mostly affecting the jaw, shoulders and nasal structure. This results in characteristic features such as dry, wrinkled skin and mottled pigmentation, reminiscent of premature ageing. Surprisingly, pubertal development proceeds normally, which distinguishes NGPS from other progeroid syndromes [183].
Individuals with NGPS usually do not present with arteriosclerosis or neurodegeneration until the third and fourth decades of life, an uncommon feature in similar diseases [184]. Despite phenotypic overlaps with HGPS, MAD and APS, NGPS is differentiated by the absence of cardiovascular complications and metabolic abnormalities. Instead, severe osteolysis involving several skeletal structures is a distinguishing feature [12] (Table 2).
NGPS exhibits a chronic progeroid phenotype due to its early onset and slow clinical progression, resulting in relatively prolonged patient survival. The clinical findings of NGPS contrast with those of classic HGPS in several respects, such as less severe growth retardation, increased height, delayed scalp hair loss and persistence of eyebrows and eyelashes [9]. Laboratory investigations have revealed vitamin D2 deficiencies and severe hypoleptinaemia, further characterising the metabolic profile of NGPS [183].
Regarding treatment, synthetic growth hormone was administered to a patient with NGPS at the age of 3 years, despite adequate endogenous growth hormone production. The treatment did not alter the natural progression of the disease [12]. Dental removal and mandibular surgery showed little efficacy in limiting mandibular resorption and may have accelerated bone deterioration [12]. Teriparatide, a recombinant human parathyroid hormone commonly used for the treatment of osteoporosis [185], was the only therapy that reduced bone loss and increased bone mineral density, although its use is not licensed. Unlike bisphosphonates, teriparatide stimulates bone remodelling, which may benefit patients with NGPS sensitive to physical stimulation. However, its long-term use is limited due to concerns about the risk of osteosarcoma [12]. Further research is needed to evaluate the safety and efficacy of aminobisphosphonates and teriparatide in NGPS. Identification of BANF1 variants associated with progeroid syndrome could lead to new therapeutics [175,186,187] and the development of NGPS mouse models would facilitate the testing of new therapies [12].

9. Discussion

The disorders that fall within the family of lipodystrophic laminopathies are defined by the presence of pathogenic variants in the same gene (LMNA), and other related genes, in addition to the presence of an impaired adipose tissue pattern. However, there is great clinical heterogeneity among these syndromes, ranging from Dunnigan disease, in which the most relevant trait is precisely adipose tissue dysfunction and lipodystrophy, to the other laminopathies affecting adipose tissue, characterised by the presence of signs of premature ageing (Table 2).
In this sense, the nuclear lamina has attracted the attention of researchers in the field of ageing considering its participation in the maintenance of the nuclear structure, DNA replication, chromatin organisation and gene expression in such a way that variants in genes encoding proteins involved in its structure and/or function can cause these premature ageing syndromes [188]. Thus, progeroid laminopathies have also become excellent sources of information for the understanding of physiological ageing [184] and debate has been opened on the extent to which the mechanisms of progeria and normal ageing may overlap. In fact, it has been speculated that defective prelamin A processing and accumulation may have a role in physiological ageing in relation with decreased expression of ZMPSTE24, although this continues to be a matter of discussion [189].
Although, to date, there is still no clear consensus on what should truly be considered premature ageing, it has recently been proposed that to tag an entity as progeroid at least 40% of ageing symptoms or signs should be present, before the typical age of onset in physiological ageing [180,184]. These ageing symptoms or signs consist not only of a series of external phenotypic features for which there is evidence that these occur as part of physiological ageing in the general population, but also of characteristics involving the internal organs that are likewise important features of ageing (osteoporosis, atherosclerosis, cancer, etc.). Thus, unlike other lipodystrophic laminopathies, although Dunnigan disease does not meet the phenotypic characteristics associated with progeroid syndromes beyond partial lipodystrophy, it is characterised by the development of certain organ abnormalities (such as non-ketotic diabetes mellitus, cardiovascular disease, etc.) at younger ages than for the general population, even leading to a lower average life expectancy [81,190]. This could raise the question of whether Dunnigan disease could be considered not only a monogenic model of lipodystrophy, but also of mild premature senescence. In this sense, whereas in the majority of lipodystrophic laminopathies associated with premature ageing, the accumulation of farnesylated prelamin A is observed [118,153,161], the involvement of prelamin A in the pathogenesis of FPLD2 should be considered with caution. Thus, while there are studies that appear to have shown prelamin A accumulation in peripheral subcutaneous adipose tissue in patients with FPLD2 [14], a more recent study, using an anti-prelamin A monoclonal antibody, did not find such accumulation in cultured fibroblasts from FPLD patients with different variants in the LMNA gene [191].
The general group of laminopathies mainly affects mesenchymal tissues (adipose tissue, muscle, bone) [10]. However, this does not occur homogeneously in all syndromes, with different disorders affecting different tissues. The presence of overlapping features is even relatively common [192]. Thus, while in the case of Dunnigan disease, only adipose tissue is affected along with muscle hypertrophy [193,194], in the case of HGPS, MADA, MADB, NGPS and APS, adipose tissue, muscle and bone are affected [195]. In addition, the pattern of lipodystrophy is not the same for each of these diseases. While Dunnigan disease, MADA and some subtypes of APS show partial lipodystrophy, in the case of HGPS, MADB, NGPS and some subtypes of APS the lipodystrophy is generalised [11]. The reason for this variable expressivity in mesenchymal tissue abnormalities, and more specifically in the fat distribution pattern of lipodystrophic laminopathies, is unknown. However, there is recent evidence that changes in the epigenome (such as in the case of genes mediated by HOTAIR or the miR-335 locus, FGF21 or LPL) can affect adipogenesis, conditioning the development and impairment of adipose tissue in a depot-specific manner [21,196]. This could help to understand the adipo-phenotype of the different subtypes of lipodystrophic laminopathies.
On the other hand, in lipodystrophy, the lack of adipocytes prevents the effective storage of lipids, which are, therefore, also accumulated ectopically, mainly in the liver and muscle [197]. This situation, along with a dysfunctional pattern of adipocytokine production, plays a relevant role in the aetopathiogenesis of insulin resistance [63]. Consequently, most lipodystrophy syndromes are characterised by a variable degree of metabolic comorbidities [197]. In addition, taking into account that the severity of insulin resistance and the toxic accumulation of fat is more severe in the generalised forms of lipodystrophy considering that adipose tissue storage capacity is non-existent by definition, patients with classic generalised lipodystrophy are associated with more severe metabolic disease [198]. However, this does not appear to be true for all types of lipodystrophic laminopathies. It is, in fact, striking that in entities in which lipodystrophy is severe, as is the case of NGPS and HGPS, which are characterised by a generalised form of lipodystrophy, the metabolic comorbidities are mild or even absent [12].
There are other multiple broad differences that make a distinction between the different lipodystrophic laminopathies, such as the age at which the first symptoms of the disease appear, ranging from the perinatal stage in the case of MADB to puberty in the case of women with FPLD2, passing through early childhood in the other lipodystrophic laminopathies [193,199]. The differences in lifespan are also striking, even within progeroid syndromes. Thus, while for the rest of the progerias average lifespan is 14 years, in the case of NGPS the patients described presented with a clearly higher life expectancy (>30 years). For this reason, this syndrome is also defined as a chronic progeria, considering the slow clinical course and longer survival [12,181]. In the same way, patients with APS and longer life expectancy (even >50 years) have also been reported [143]. This is consistent with the complex genotype-phenotype associations and clinical heterogeneity of laminopathies [11].
Lipodystrophic laminopathies are considered to be rare disorders, or even ultra-rare disorders based on the prevalence rate of less than 1 per 50,000 individuals defined by the European Union for ultra-rare diseases (https://eur-lex.europa.eu/legal-content/EN/TXT/?uri=CELEX%3A32014R0536&qid=1720375578947, accessed on 7 July 2024). However, it is difficult to speak of prevalence in a rare disease. In fact, for many of these ultra-rare disorders, prevalence cannot be established taking into account the low number of cases described in the literature to date. In addition, while premature ageing syndromes are more clinically recognisable and ultimately more easily diagnosed, in the case of Dunnigan disease the diagnosis is more intricate, especially in men in whom the phenotype is less apparent [41,42]. Furthermore, the clinical heterogeneity previously described even within the same syndrome may also contribute to delays in diagnosis or even misdiagnosis. These factors are of special relevance considering that the initial diagnosis of these diseases is merely clinical. Thus, the estimation of prevalence for these syndromes will not only be overshadowed by the small number of reported cases, but also underestimated due to the late diagnosis or underdiagnosis in some cases. The presence of specialised reference units with multidisciplinary teams for the diagnosis and management of these rare disorders is essential, along with international research cooperation to help carry out studies with larger sample sizes and present a more objective picture of these syndromes.
In conclusion, lipodystrophic laminopathies are ultra-rare disorders due to pathogenic variants in the same gene (LMNA), and other related genes, that condition alterations in lamin A processing and, consequently, adipose tissue impairment along with diverse clinical manifestations and comorbidities. Despite the similarities in their aetiopathogenesis, they prove to be individual entities with complex genotype-phenotype associations and great clinical heterogeneity as one of their most fascinating traits. Therefore, there are still many questions to be resolved and, although in recent years new therapies have been developed primarily aimed at the management of their comorbidities, there is still much to do. Thus, international networks should be established in order to better understand these disorders.

10. Research Agenda

-
There is a need to establish international cooperation and networks through the International Clinical Working Groups of the European Consortium of Lipodystrophies (ECLip) in order to address the real lipodystrophic laminopathies and create registries that bring together relevant information about these ultra-rare pathologies.
-
There is also a need to delve deeper into what should really be considered a progeroid syndrome while trying to gather scientific evidence based on consensus among worldwide experts.
-
Increasing knowledge of the molecular basis of lipodystrophic laminopathies will make it possible, in the future, to find new specific therapies to shed light on what until now are life-threatening or fatal disorders.
-
Attention should be focused on the field of epigenomics and the possibility of modulating clinical expressivity through reversible epigenetic markers. In this sense, epigenetic studies on whole adipose tissue and preadipocytes, including the already-known miR335 but also HOTAIR involvement and other epigenetic marks, could improve understanding of the big phenotypic differences (when comparing FPLD vs. progeroid syndromes), or mild phenotypic differences in FPLD2 pending the specific LMNA variants.

Author Contributions

Conceptualisation, A.F.-P. and D.A.-V.; writing—original draft preparation, E.J.D.-L. and A.F.-P.; writing—review and editing, D.A.-V., S.S.-I., A.I.C., S.C.-G. and T.P.-M.; supervision, D.A.-V. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Instituto de Salud Carlos III (grant PI22/00514) and co-funded by the European Union and an intramural grant from the Xunta de Galicia, ED431B 2020/37. A.F.-P. is funded by a Juan Rodes research contract (JR23/00042) from the Instituto de Salud Carlos III (ISCIII), co-financed by the ERDF. S.S.-I. was awarded a Research Fellowship by the Asociación Española de Familiares y Afectados de Lipodistrofias (AELIP).

Institutional Review Board Statement

The study was conducted in accordance with the Declaration of Helsinki and approved by the ethics review panel of the Red Gallega de Comités de Ética de la Investigación (approval code 2022/383, date of approval 23 November 2022).

Informed Consent Statement

Informed consent was obtained from all subjects for the publication of their pictures and clinical, imaging and molecular information.

Data Availability Statement

No new data were analysed in this study. Data sharing is not applicable to this article.

Acknowledgments

We are indebted to the patients of this study for their collaboration.

Conflicts of Interest

D.A.-V. has received fees from Amryt Pharmaceuticals and Regeneron Pharmaceuticals for scientific advice, travel, conference registration and research grants. The rest of the authors declare no conflicts of interest.

References

  1. Bidault, G.; Vatier, C.; Capeau, J.; Vigouroux, C.; Béréziat, V. LMNA-Linked Lipodystrophies: From Altered Fat Distribution to Cellular Alterations. Biochem. Soc. Trans. 2011, 39, 1752–1757. [Google Scholar] [CrossRef] [PubMed]
  2. Perepelina, K.; Klauzen, P.; Kostareva, A.; Malashicheva, A. Tissue-Specific Influence of Lamin A Mutations on Notch Signaling and Osteogenic Phenotype of Primary Human Mesenchymal Cells. Cells 2019, 8, 266. [Google Scholar] [CrossRef] [PubMed]
  3. Gruenbaum, Y.; Goldman, R.D.; Meyuhas, R.; Mills, E.; Margalit, A.; Fridkin, A.; Dayani, Y.; Prokocimer, M.; Enosh, A. The Nuclear Lamina and Its Functions in the Nucleus. Int. Rev. Cytol. 2003, 226, 1–62. [Google Scholar] [CrossRef]
  4. Jacob, K.N.; Garg, A. Laminopathies: Multisystem Dystrophy Syndromes. Mol. Genet. Metab. 2006, 87, 289–302. [Google Scholar] [CrossRef] [PubMed]
  5. Dauer, W.T.; Worman, H.J. The Nuclear Envelope as a Signaling Node in Development and Disease. Dev. Cell. 2009, 17, 626–638. [Google Scholar] [CrossRef]
  6. Garg, A. Acquired and Inherited Lipodystrophies. N. Engl. J. Med. 2004, 350, 1220–1234. [Google Scholar] [CrossRef]
  7. Capanni, C.; Cenni, V.; Haraguchi, T.; Squarzoni, S.; Schüchner, S.; Ogris, E.; Novelli, G.; Maraldi, N.; Lattanzi, G. Lamin A Precursor Induces Barrier-to-Autointegration Factor Nuclear Localization. Celli Cycle 2010, 9, 2600–2610. [Google Scholar] [CrossRef]
  8. Cenni, V.; Capanni, C.; Mattioli, E.; Columbaro, M.; Wehnert, M.; Ortolani, M.; Fini, M.; Novelli, G.; Bertacchini, J.; Maraldi, N.M.; et al. Rapamycin Treatment of Mandibuloacral Dysplasia Cells Rescues Localization of Chromatin-Associated Proteins and Cell Cycle Dynamics. Aging 2014, 6, 755–770. [Google Scholar] [CrossRef]
  9. Osorio, F.G.; Ugalde, A.P.; Mariño, G.; Puente, X.S.; Freije, J.M.P.; López-Otín, C. Cell Autonomous and Systemic Factors in Progeria Development. Biochem. Soc. Trans. 2011, 39, 1710–1714. [Google Scholar] [CrossRef]
  10. Broers, J.L.V.; Ramaekers, F.C.S.; Bonne, G.; Yaou, R.B.; Hutchison, C.J. Nuclear Lamins: Laminopathies and Their Role in Premature Ageing. Physiol. Rev. 2006, 86, 967–1008. [Google Scholar] [CrossRef]
  11. Guillín-Amarelle, C.; Fernández-Pombo, A.; Sánchez-Iglesias, S.; Araújo-Vilar, D. Lipodystrophic Laminopathies: Diagnostic Clues. Nucleus 2018, 9, 249–260. [Google Scholar] [CrossRef]
  12. Cabanillas, R.; Cadiñanos, J.; Villameytide, J.A.F.; Pérez, M.; Longo, J.; Richard, J.M.; Alvarez, R.; Durán, N.S.; Illán, R.; González, D.J.; et al. Néstor-Guillermo Progeria Syndrome: A Novel Premature Aging Condition with Early Onset and Chronic Development Caused by BANF1 Mutations. Am. J. Med. Genet. A 2011, 155, 2617–2625. [Google Scholar] [CrossRef] [PubMed]
  13. Davies, B.S.J.; Barnes, R.H.; Tu, Y.; Ren, S.; Andres, D.A.; Spielmann, H.P.; Lammerding, J.; Wang, Y.; Young, S.G.; Fong, L.G. An Accumulation of Non-Farnesylated Prelamin A Causes Cardiomyopathy but Not Progeria. Hum. Mol. Genet. 2010, 19, 2682–2694. [Google Scholar] [CrossRef]
  14. Araújo-Vilar, D.; Lattanzi, G.; González-Méndez, B.; Costa-Freitas, A.T.; Prieto, D.; Columbaro, M.; Mattioli, E.; Victoria, B.; Martínez-Sánchez, N.; Ramazanova, A.; et al. Site-Dependent Differences in Both Prelamin A and Adipogenic Genes in Subcutaneous Adipose Tissue of Patients with Type 2 Familial Partial Lipodystrophy. J. Med. Genet. 2009, 46, 40–48. [Google Scholar] [CrossRef]
  15. Candelario, J.; Borrego, S.; Reddy, S.; Comai, L. Accumulation of Distinct Prelamin A Variants in Human Diploid Fibroblasts Differentially Affects Cell Homeostasis. Exp. Cell. Res. 2011, 317, 319–329. [Google Scholar] [CrossRef] [PubMed]
  16. Davies, B.S.; Coffinier, C.; Yang, S.H.; Barnes, R.H.; Jung, H.-J.; Young, S.G.; Fong, L.G. Investigating the Purpose of Prelamin A Processing. Nucleus 2011, 2, 4–9. [Google Scholar] [CrossRef] [PubMed]
  17. Dechat, T.; Pfleghaar, K.; Sengupta, K.; Shimi, T.; Shumaker, D.K.; Solimando, L.; Goldman, R.D. Nuclear Lamins: Major Factors in the Structural Organization and Function of the Nucleus and Chromatin. Genes. Dev. 2008, 22, 832–853. [Google Scholar] [CrossRef] [PubMed]
  18. Gonzalez-Suarez, I.; Gonzalo, S. Nurturing the Genome: A-Type Lamins Preserve Genomic Stability. Nucleus 2010, 1, 129–135. [Google Scholar] [CrossRef]
  19. Liu, B.; Wang, J.; Chan, K.M.; Tjia, W.M.; Deng, W.; Guan, X.; Huang, J.; Li, K.M.; Chau, P.Y.; Chen, D.J.; et al. Genomic Instability in Laminopathy-Based Premature Aging. Nat. Med. 2005, 11, 780–785. [Google Scholar] [CrossRef]
  20. Muralikrishna, B.; Dhawan, J.; Rangaraj, N.; Parnaik, V.K. Distinct Changes in Intranuclear Lamin A/C Organization during Myoblast Differentiation. J. Cell. Sci. 2001, 114, 4001–4011. [Google Scholar] [CrossRef]
  21. Oldenburg, A.; Briand, N.; Sørensen, A.L.; Cahyani, I.; Shah, A.; Moskaug, J.Ø.; Collas, P. A Lipodystrophy-Causing Lamin A Mutant Alters Conformation and Epigenetic Regulation of the Anti-Adipogenic MIR335 Locus. J. Cell. Biol. 2017, 216, 2731–2743. [Google Scholar] [CrossRef] [PubMed]
  22. Osmanagic-Myers, S.; Dechat, T.; Foisner, R. Lamins at the Crossroads of Mechanosignaling. Genes. Dev. 2015, 29, 225–237. [Google Scholar] [CrossRef] [PubMed]
  23. Pellegrini, C.; Columbaro, M.; Schena, E.; Prencipe, S.; Andrenacci, D.; Iozzo, P.; Angela Guzzardi, M.; Capanni, C.; Mattioli, E.; Loi, M.; et al. Altered Adipocyte Differentiation and Unbalanced Autophagy in Type 2 Familial Partial Lipodystrophy: An in Vitro and in Vivo Study of Adipose Tissue Browning. Exp. Mol. Med. 2019, 51, 1–17. [Google Scholar] [CrossRef]
  24. Varela, I.; Cadiñanos, J.; Pendás, A.M.; Gutiérrez-Fernández, A.; Folgueras, A.R.; Sánchez, L.M.; Zhou, Z.; Rodríguez, F.J.; Stewart, C.L.; Vega, J.A.; et al. Accelerated Ageing in Mice Deficient in Zmpste24 Protease Is Linked to P53 Signalling Activation. Nature 2005, 437, 564–568. [Google Scholar] [CrossRef] [PubMed]
  25. Vigouroux, C.; Auclair, M.; Dubosclard, E.; Pouchelet, M.; Capeau, J.; Courvalin, J.C.; Buendia, B. Nuclear Envelope Disorganization in Fibroblasts from Lipodystrophic Patients with Heterozygous R482Q/W Mutations in the Lamin A/C Gene. J Cell. Sci. 2001, 114, 4459–4468. [Google Scholar] [CrossRef]
  26. Guénantin, A.C.; Briand, N.; Bidault, G.; Afonso, P.; Béréziat, V.; Vatier, C.; Lascols, O.; Caron-Debarle, M.; Capeau, J.; Vigouroux, C. Nuclear Envelope-Related Lipodystrophies. Semin. Cell. Dev. Biol. 2014, 29, 148–157. [Google Scholar] [CrossRef]
  27. Lloyd, D.J.; Trembath, R.C.; Shackleton, S. A Novel Interaction between Lamin A and SREBP1: Implications for Partial Lipodystrophy and Other Laminopathies. Hum. Mol. Genet. 2002, 11, 769–777. [Google Scholar] [CrossRef]
  28. Capanni, C.; Squarzoni, S.; Cenni, V.; D’Apice, M.R.; Gambineri, A.; Novelli, G.; Wehnert, M.; Pasquali, R.; Maraldi, N.M.; Lattanzi, G. Familial Partial Lipodystrophy, Mandibuloacral Dysplasia and Restrictive Dermopathy Feature Barrier-to-Autointegration Factor (BAF) Nuclear Redistribution. Celli Cycle 2012, 11, 3568–3577. [Google Scholar] [CrossRef]
  29. Johnson, B.R.; Nitta, R.T.; Frock, R.L.; Mounkes, L.; Barbie, D.A.; Stewart, C.L.; Harlow, E.; Kennedy, B.K. A-Type Lamins Regulate Retinoblastoma Protein Function by Promoting Subnuclear Localization and Preventing Proteasomal Degradation. Proc. Natl. Acad. Sci. USA 2004, 101, 9677–9682. [Google Scholar] [CrossRef]
  30. Araújo-Vilar, D.; Victoria, B.; González-Méndez, B.; Barreiro, F.; Fernández-Rodríguez, B.; Cereijo, R.; Gallego-Escuredo, J.M.; Villarroya, F.; Pañeda-Menéndez, A. Histological and Molecular Features of Lipomatous and Nonlipomatous Adipose Tissue in Familial Partial Lipodystrophy Caused by LMNA Mutations. Clin. Endocrinol. 2012, 76, 816–824. [Google Scholar] [CrossRef]
  31. Mateos, J.; Landeira-Abia, A.; Fafián-Labora, J.A.; Fernández-Pernas, P.; Lesende-Rodríguez, I.; Fernández-Puente, P.; Fernández-Moreno, M.; Delmiro, A.; Martín, M.A.; Blanco, F.J.; et al. iTRAQ-Based Analysis of Progerin Expression Reveals Mitochondrial Dysfunction, Reactive Oxygen Species Accumulation and Altered Proteostasis. Stem. Cell. Res. Ther. 2015, 6, 119. [Google Scholar] [CrossRef] [PubMed]
  32. Afonso, P.; Auclair, M.; Boccara, F.; Vantyghem, M.-C.; Katlama, C.; Capeau, J.; Vigouroux, C.; Caron-Debarle, M. LMNA Mutations Resulting in Lipodystrophy and HIV Protease Inhibitors Trigger Vascular Smooth Muscle Cell Senescence and Calcification: Role of ZMPSTE24 Downregulation. Atherosclerosis 2016, 245, 200–211. [Google Scholar] [CrossRef]
  33. Gabriel, D.; Roedl, D.; Gordon, L.B.; Djabali, K. Sulforaphane Enhances Progerin Clearance in Hutchinson-Gilford Progeria Fibroblasts. Aging. Cell. 2015, 14, 78–91. [Google Scholar] [CrossRef]
  34. Gruenbaum, Y.; Foisner, R. Lamins: Nuclear Intermediate Filament Proteins with Fundamental Functions in Nuclear Mechanics and Genome Regulation. Annu. Rev. Biochem. 2015, 84, 131–164. [Google Scholar] [CrossRef]
  35. Dittmer, T.A.; Misteli, T. The Lamin Protein Family. Genome. Biol. 2011, 12, 222. [Google Scholar] [CrossRef]
  36. Håkelien, A.-M.; Delbarre, E.; Gaustad, K.G.; Buendia, B.; Collas, P. Expression of the Myodystrophic R453W Mutation of Lamin A in C2C12 Myoblasts Causes Promoter-Specific and Global Epigenetic Defects. Exp. Cell. Res. 2008, 314, 1869–1880. [Google Scholar] [CrossRef]
  37. Arancio, W.; Pizzolanti, G.; Genovese, S.I.; Pitrone, M.; Giordano, C. Epigenetic Involvement in Hutchinson-Gilford Progeria Syndrome: A Mini-Review. Gerontology 2014, 60, 197–203. [Google Scholar] [CrossRef]
  38. Davidson, M.B.; Young, R.T. Metabolic Studies in Familial Partial Lipodystrophy of the Lower Trunk and Extremities. Diabetologia 1975, 11, 561–568. [Google Scholar] [CrossRef] [PubMed]
  39. Drac, H.; Madej-Pilarczyk, A.; Gospodarczyk-Szot, K.; Gaweł, M.; Kwieciński, H.; Hausmanowa-Petrusewicz, I. Familial Partial Lipodystrophy Associated with the Heterozygous LMNA Mutation 1445G>A (Arg482Gln) in a Polish Family. Neurol. Neurochir. Pol. 2010, 44, 291–296. [Google Scholar] [CrossRef] [PubMed]
  40. Köbberling, J.; Dunnigan, M.G. Familial Partial Lipodystrophy: Two Types of an X Linked Dominant Syndrome, Lethal in the Hemizygous State. J. Med. Genet. 1986, 23, 120–127. [Google Scholar] [CrossRef]
  41. Vigouroux, C.; Magré, J.; Vantyghem, M.C.; Bourut, C.; Lascols, O.; Shackleton, S.; Lloyd, D.J.; Guerci, B.; Padova, G.; Valensi, P.; et al. Lamin A/C Gene: Sex-Determined Expression of Mutations in Dunnigan-Type Familial Partial Lipodystrophy and Absence of Coding Mutations in Congenital and Acquired Generalized Lipoatrophy. Diabetes 2000, 49, 1958–1962. [Google Scholar] [CrossRef] [PubMed]
  42. Akinci, B.; Onay, H.; Demir, T.; Savas-Erdeve, Ş.; Gen, R.; Simsir, I.Y.; Keskin, F.E.; Erturk, M.S.; Uzum, A.K.; Yaylali, G.F.; et al. Clinical Presentations, Metabolic Abnormalities and End-Organ Complications in Patients with Familial Partial Lipodystrophy. Metabolism 2017, 72, 109–119. [Google Scholar] [CrossRef] [PubMed]
  43. Araújo-Vilar, D.; Fernández-Pombo, A.; Victoria, B.; Mosquera-Orgueira, A.; Cobelo-Gómez, S.; Castro-Pais, A.; Hermida-Ameijeiras, Á.; Loidi, L.; Sánchez-Iglesias, S. Variable Expressivity and Allelic Heterogeneity in Type 2 Familial Partial Lipodystrophy: The p.(Thr528Met) LMNA Variant. J. Clin. Med. 2021, 10, 1497. [Google Scholar] [CrossRef] [PubMed]
  44. de Andrade, N.X.S.; Adiyaman, S.C.; Yuksel, B.D.; Ferrari, C.T.; Eldin, A.J.; Saydam, B.O.; Altay, C.; Sharma, P.; Bhave, N.; Little, A.; et al. Unusual presentations of LMNA-associated lipodystrophy with complex phenotypes and generalized fat loss: When the genetic diagnosis uncovers novel features. AACE Clin. Case. Rep. 2020, 6, e79–e85. [Google Scholar] [CrossRef]
  45. Decaudain, A.; Vantyghem, M.-C.; Guerci, B.; Hécart, A.-C.; Auclair, M.; Reznik, Y.; Narbonne, H.; Ducluzeau, P.-H.; Donadille, B.; Lebbé, C.; et al. New Metabolic Phenotypes in Laminopathies: LMNA Mutations in Patients with Severe Metabolic Syndrome. J. Clin. Endocrinol. Metab. 2007, 92, 4835–4844. [Google Scholar] [CrossRef]
  46. Garg, A. Gender Differences in the Prevalence of Metabolic Complications in Familial Partial Lipodystrophy (Dunnigan Variety). J. Clin. Endocrinol. Metab. 2000, 85, 1776–1782. [Google Scholar] [CrossRef]
  47. Montenegro, R.M.; Costa-Riquetto, A.D.; Fernandes, V.O.; Montenegro, A.P.D.R.; de Santana, L.S.; de Lima Jorge, A.A.; de Azevedo Souza Karbage, L.B.; Aguiar, L.B.; Carvalho, F.H.C.; Teles, M.G.; et al. Homozygous and Heterozygous Nuclear Lamin A p.R582C Mutation: Different Lipodystrophic Phenotypes in the Same Kindred. Front. Endocrinol. 2018, 9, 458. [Google Scholar] [CrossRef]
  48. Mory, P.B.; Crispim, F.; Freire, M.B.S.; Salles, J.E.N.; Valério, C.M.; Godoy-Matos, A.F.; Dib, S.A.; Moisés, R.S. Phenotypic Diversity in Patients with Lipodystrophy Associated with LMNA Mutations. Eur. J. Endocrinol. 2012, 167, 423–431. [Google Scholar] [CrossRef]
  49. Muschke, P.; Kölsch, U.; Jakubiczka, S.; Wieland, I.; Brune, T.; Wieacker, P. The Heterozygous LMNA Mutation p.R471G Causes a Variable Phenotype with Features of Two Types of Familial Partial Lipodystrophy. Am. J. Med. Genet. A 2007, 143, 2810–2814. [Google Scholar] [CrossRef]
  50. Resende, A.T.P.; Martins, C.S.; Bueno, A.C.; Moreira, A.C.; Foss-Freitas, M.C.; de Castro, M. Phenotypic Diversity and Glucocorticoid Sensitivity in Patients with Familial Partial Lipodystrophy Type 2. Clin. Endocrinol. 2019, 91, 94–103. [Google Scholar] [CrossRef]
  51. Sorkina, E.L.; Kalashnikova, M.F.; Melnichenko, G.A.; Tyulpakov, A.N. Familial partial lipodystrophy (Dunnigan syndrome) due to LMNA gene mutation: The first description of its clinical case in Russia. Ter. Arkhiv 2015, 87, 83–87. [Google Scholar] [CrossRef]
  52. Gonzaga-Jauregui, C.; Ge, W.; Staples, J.; Van Hout, C.; Yadav, A.; Colonie, R.; Leader, J.B.; Kirchner, H.L.; Murray, M.F.; Reid, J.G.; et al. Clinical and Molecular Prevalence of Lipodystrophy in an Unascertained Large Clinical Care Cohort. Diabetes 2020, 69, 249–258. [Google Scholar] [CrossRef]
  53. Spuler, S.; Kalbhenn, T.; Zabojszcza, J.; van Landeghem, F.K.H.; Ludtke, A.; Wenzel, K.; Koehnlein, M.; Schuelke, M.; Lüdemann, L.; Schmidt, H.H. Muscle and Nerve Pathology in Dunnigan Familial Partial Lipodystrophy. Neurology 2007, 68, 677–683. [Google Scholar] [CrossRef] [PubMed]
  54. Akinci, G.; Topaloglu, H.; Demir, T.; Danyeli, A.E.; Talim, B.; Keskin, F.E.; Kadioglu, P.; Talip, E.; Altay, C.; Yaylali, G.F.; et al. Clinical Spectra of Neuromuscular Manifestations in Patients with Lipodystrophy: A Multicenter Study. Neuromuscul. Disord. 2017, 27, 923–930. [Google Scholar] [CrossRef] [PubMed]
  55. Burn, J.; Baraitser, M. Partial Lipoatrophy with Insulin Resistant Diabetes and Hyperlipidaemia (Dunnigan Syndrome). J. Med. Genet. 1986, 23, 128–130. [Google Scholar] [CrossRef]
  56. Vantyghem, M.C.; Pigny, P.; Maurage, C.A.; Rouaix-Emery, N.; Stojkovic, T.; Cuisset, J.M.; Millaire, A.; Lascols, O.; Vermersch, P.; Wemeau, J.L.; et al. Patients with Familial Partial Lipodystrophy of the Dunnigan Type Due to a LMNA R482W Mutation Show Muscular and Cardiac Abnormalities. J. Clin. Endocrinol. Metab. 2004, 89, 5337–5346. [Google Scholar] [CrossRef]
  57. Morel, C.F.; Thomas, M.A.; Cao, H.; O’Neil, C.H.; Pickering, J.G.; Foulkes, W.D.; Hegele, R.A. A LMNA Splicing Mutation in Two Sisters with Severe Dunnigan-Type Familial Partial Lipodystrophy Type 2. J. Clin. Endocrinol. Metab. 2006, 91, 2689–2695. [Google Scholar] [CrossRef]
  58. Treiber, G.; Flaus Furmaniuk, A.; Guilleux, A.; Medjane, S.; Bonfanti, O.; Schneebeli, S.; Bernard, C.; Le-Moullec, N.; Bakiri, F.; Pholsena, M.; et al. A Recurrent Familial Partial Lipodystrophy Due to a Monoallelic or Biallelic LMNA Founder Variant Highlights the Multifaceted Cardiac Manifestations of Metabolic Laminopathies. Eur. J. Endocrinol. 2021, 185, 453–462. [Google Scholar] [CrossRef] [PubMed]
  59. Vaidya, R.A.; Vaidya, A.D.B.; Talwalkar, S.C.; Mehtalia, S.D.; Shringi, M.S.; Pandey, S.N.; Shah, S.J.; Godse, C.; Joshi, J.V.; Sheth, J.; et al. Clinical, Endocrine and Metabolic Studies in the Kindred of Familial Partial Lipodystrophy--a Syndrome of Insulin Resistance. J. Assoc. Physicians India 2002, 50, 773–776. [Google Scholar]
  60. Jeru, I.; Vatier, C.; Vantyghem, M.-C.; Lascols, O.; Vigouroux, C. LMNA-Associated Partial Lipodystrophy: Anticipation of Metabolic Complications. J. Med. Genet. 2017, 54, 413–416. [Google Scholar] [CrossRef]
  61. Patni, N.; Li, X.; Adams-Huet, B.; Vasandani, C.; Gomez-Diaz, R.A.; Garg, A. Regional Body Fat Changes and Metabolic Complications in Children With Dunnigan Lipodystrophy-Causing LMNA Variants. J. Clin. Endocrinol. Metab. 2019, 104, 1099–1108. [Google Scholar] [CrossRef] [PubMed]
  62. Zhong, Z.X.; Harris, J.; Wilber, E.; Gorman, S.; Savage, D.B.; O’Rahilly, S.; Stears, A.; Williams, R.M. Describing the Natural History of Clinical, Biochemical and Radiological Outcomes of Children with Familial Partial Lipodystrophy Type 2 (FPLD2) from the United Kingdom: A Retrospective Case Series. Clin. Endocrinol. 2022, 97, 755–762. [Google Scholar] [CrossRef] [PubMed]
  63. Huang-Doran, I.; Sleigh, A.; Rochford, J.J.; O’Rahilly, S.; Savage, D.B. Lipodystrophy: Metabolic Insights from a Rare Disorder. J. Endocrinol. 2010, 207, 245–255. [Google Scholar] [CrossRef] [PubMed]
  64. Cao, H.; Hegele, R.A. Nuclear Lamin A/C R482Q Mutation in Canadian Kindreds with Dunnigan-Type Familial Partial Lipodystrophy. Hum. Mol. Genet. 2000, 9, 109–112. [Google Scholar] [CrossRef]
  65. Haque, W.A.; Oral, E.A.; Dietz, K.; Bowcock, A.M.; Agarwal, A.K.; Garg, A. Risk Factors for Diabetes in Familial Partial Lipodystrophy, Dunnigan Variety. Diabetes Care 2003, 26, 1350–1355. [Google Scholar] [CrossRef]
  66. Joy, T.; Kennedy, B.A.; Al-Attar, S.; Rutt, B.K.; Hegele, R.A. Predicting Abdominal Adipose Tissue among Women with Familial Partial Lipodystrophy. Metabolism 2009, 58, 828–834. [Google Scholar] [CrossRef]
  67. Lazarte, J.; Wang, J.; McIntyre, A.D.; Hegele, R.A. Prevalence of Severe Hypertriglyceridemia and Pancreatitis in Familial Partial Lipodystrophy Type 2. J. Clin. Lipidol. 2021, 15, 653–657. [Google Scholar] [CrossRef]
  68. Treiber, G.; Guilleux, A.; Huynh, K.; Bonfanti, O.; Flaus-Furmaniuk, A.; Couret, D.; Mellet, N.; Bernard, C.; Le-Moullec, N.; Doray, B.; et al. Lipoatrophic Diabetes in Familial Partial Lipodystrophy Type 2: From Insulin Resistance to Diabetes. Diabetes. Metab. 2023, 49, 101409. [Google Scholar] [CrossRef]
  69. Haque, W.A.; Vuitch, F.; Garg, A. Post-Mortem Findings in Familial Partial Lipodystrophy, Dunnigan Variety. Diabet. Med. 2002, 19, 1022–1025. [Google Scholar] [CrossRef]
  70. Foss-Freitas, M.C.; Ferraz, R.C.; Monteiro, L.Z.; Gomes, P.M.; Iwakura, R.; de Freitas, L.C.C.; Foss, M.C. Endoplasmic Reticulum Stress Activation in Adipose Tissue Induces Metabolic Syndrome in Individuals with Familial Partial Lipodystrophy of the Dunnigan Type. Diabetol. Metab. Syndr. 2018, 10, 6. [Google Scholar] [CrossRef]
  71. Hegele, R.A.; Kraw, M.E.; Ban, M.R.; Miskie, B.A.; Huff, M.W.; Cao, H. Elevated Serum C-Reactive Protein and Free Fatty Acids among Nondiabetic Carriers of Missense Mutations in the Gene Encoding Lamin A/C (LMNA) with Partial Lipodystrophy. Arterioscler. Thromb. Vasc. Biol. 2003, 23, 111–116. [Google Scholar] [CrossRef]
  72. Kutbay, N.O.; Yurekli, B.S.; Onay, H.; Altay, C.T.; Atik, T.; Hekimsoy, Z.; Saygili, F.; Akinci, B. A Case of Familial Partial Lipodystrophy Caused by a Novel Lamin A/C (LMNA) Mutation in Exon 1 (D47N). Eur. J. Intern. Med. 2016, 29, 37–39. [Google Scholar] [CrossRef] [PubMed]
  73. Speckman, R.A.; Garg, A.; Du, F.; Bennett, L.; Veile, R.; Arioglu, E.; Taylor, S.I.; Lovett, M.; Bowcock, A.M. Mutational and Haplotype Analyses of Families with Familial Partial Lipodystrophy (Dunnigan Variety) Reveal Recurrent Missense Mutations in the Globular C-Terminal Domain of Lamin A/C. Am. J. Hum. Genet. 2000, 66, 1192–1198. [Google Scholar] [CrossRef] [PubMed]
  74. Vasandani, C.; Li, X.; Sekizkardes, H.; Brown, R.J.; Garg, A. Phenotypic Differences Among Familial Partial Lipodystrophy Due to LMNA or PPARG Variants. J. Endocr. Soc. 2022, 6, bvac155. [Google Scholar] [CrossRef] [PubMed]
  75. Akinci, B.; Unlu, S.M.; Celik, A.; Simsir, I.Y.; Sen, S.; Nur, B.; Keskin, F.E.; Ozgen Saydam, B.; Kutbay Ozdemir, N.; Sarer Yurekli, B.; et al. Renal Complications of Lipodystrophy: A Closer Look at the Natural History of Kidney Disease. Clin. Endocrinol. 2018, 89, 65–75. [Google Scholar] [CrossRef]
  76. Ajluni, N.; Meral, R.; Neidert, A.H.; Brady, G.F.; Buras, E.; McKenna, B.; DiPaola, F.; Chenevert, T.L.; Horowitz, J.F.; Buggs-Saxton, C.; et al. Spectrum of Disease Associated with Partial Lipodystrophy: Lessons from a Trial Cohort. Clin. Endocrinol. 2017, 86, 698–707. [Google Scholar] [CrossRef]
  77. Araújo-Vilar, D.; Lado-Abeal, J.; Palos-Paz, F.; Lattanzi, G.; Bandín, M.A.; Bellido, D.; Domínguez-Gerpe, L.; Calvo, C.; Pérez, O.; Ramazanova, A.; et al. A Novel Phenotypic Expression Associated with a New Mutation in LMNA Gene, Characterized by Partial Lipodystrophy, Insulin Resistance, Aortic Stenosis and Hypertrophic Cardiomyopathy. Clin. Endocrinol. 2008, 69, 61–68. [Google Scholar] [CrossRef] [PubMed]
  78. Garg, A.; Speckman, R.A.; Bowcock, A.M. Multisystem Dystrophy Syndrome Due to Novel Missense Mutations in the Amino-Terminal Head and Alpha-Helical Rod Domains of the Lamin A/C Gene. Am. J. Med. 2002, 112, 549–555. [Google Scholar] [CrossRef]
  79. Subramanyam, L.; Simha, V.; Garg, A. Overlapping Syndrome with Familial Partial Lipodystrophy, Dunnigan Variety and Cardiomyopathy Due to Amino-Terminal Heterozygous Missense Lamin A/C Mutations. Clin. Genet. 2010, 78, 66–73. [Google Scholar] [CrossRef]
  80. Bidault, G.; Garcia, M.; Vantyghem, M.-C.; Ducluzeau, P.-H.; Morichon, R.; Thiyagarajah, K.; Moritz, S.; Capeau, J.; Vigouroux, C.; Béréziat, V. Lipodystrophy-Linked LMNA p.R482W Mutation Induces Clinical Early Atherosclerosis and In Vitro Endothelial Dysfunction. Arterioscler. Thromb. Vasc. Biol. 2013, 33, 2162–2171. [Google Scholar] [CrossRef]
  81. Hegele, R.A. Premature Atherosclerosis Associated with Monogenic Insulin Resistance. Circulation 2001, 103, 2225–2229. [Google Scholar] [CrossRef] [PubMed]
  82. Weterings, A.A.W.; van Rijsingen, I.A.W.; Plomp, A.S.; Zwinderman, A.H.; Lekanne Deprez, R.H.; Mannens, M.M.; van den Bergh Weerman, M.A.; van der Wal, A.C.; Pinto-Sietsma, S.J. A Novel Lamin A/C Mutation in a Dutch Family with Premature Atherosclerosis. Atherosclerosis 2013, 229, 169–173. [Google Scholar] [CrossRef]
  83. Eldin, A.J.; Akinci, B.; da Rocha, A.M.; Meral, R.; Simsir, I.Y.; Adiyaman, S.C.; Ozpelit, E.; Bhave, N.; Gen, R.; Yurekli, B.; et al. Cardiac Phenotype in Familial Partial Lipodystrophy. Clin. Endocrinol. 2021, 94, 1043–1053. [Google Scholar] [CrossRef] [PubMed]
  84. Gambineri, A.; Semple, R.K.; Forlani, G.; Genghini, S.; Grassi, I.; Hyden, C.S.S.; Pagotto, U.; O’Rahilly, S.; Pasquali, R. Monogenic Polycystic Ovary Syndrome Due to a Mutation in the Lamin A/C Gene Is Sensitive to Thiazolidinediones but Not to Metformin. Eur. J. Endocrinol. 2008, 159, 347–353. [Google Scholar] [CrossRef]
  85. Hegele, R.A. Lessons from Human Mutations in PPARgamma. Int. J. Obes. 2005, 29 (Suppl. S1), S31–S35. [Google Scholar] [CrossRef] [PubMed]
  86. Joy, T.R.; Hegele, R.A. Prevalence of Reproductive Abnormalities among Women with Familial Partial Lipodystrophy. Endocr. Pract. 2008, 14, 1126–1132. [Google Scholar] [CrossRef]
  87. Vantyghem, M.C.; Vincent-Desplanques, D.; Defrance-Faivre, F.; Capeau, J.; Fermon, C.; Valat, A.S.; Lascols, O.; Hecart, A.C.; Pigny, P.; Delemer, B.; et al. Fertility and Obstetrical Complications in Women with LMNA-Related Familial Partial Lipodystrophy. J. Clin. Endocrinol. Metab. 2008, 93, 2223–2229. [Google Scholar] [CrossRef]
  88. Al-Shali, K.; Cao, H.; Knoers, N.; Hermus, A.R.; Tack, C.J.; Hegele, R.A. A Single-Base Mutation in the Peroxisome Proliferator-Activated Receptor Gamma4 Promoter Associated with Altered in Vitro Expression and Partial Lipodystrophy. J. Clin. Endocrinol. Metab. 2004, 89, 5655–5660. [Google Scholar] [CrossRef]
  89. Auclair, M.; Vigouroux, C.; Boccara, F.; Capel, E.; Vigeral, C.; Guerci, B.; Lascols, O.; Capeau, J.; Caron-Debarle, M. Peroxisome Proliferator-Activated Receptor-γ Mutations Responsible for Lipodystrophy with Severe Hypertension Activate the Cellular Renin-Angiotensin System. Arterioscler. Thromb. Vasc. Biol. 2013, 33, 829–838. [Google Scholar] [CrossRef]
  90. Berger, J.R.; Oral, E.A.; Taylor, S.I. Familial Lipodystrophy Associated with Neurodegeneration and Congenital Cataracts. Neurology 2002, 58, 43–47. [Google Scholar] [CrossRef]
  91. Carboni, N.; Brancati, F.; Cocco, E.; Solla, E.; D’Apice, M.R.; Mateddu, A.; McIntyre, A.; Fadda, E.; Mura, M.; Lattanzi, G.; et al. Partial Lipodystrophy Associated with Muscular Dystrophy of Unknown Genetic Origin. Muscle Nerve 2014, 49, 928–930. [Google Scholar] [CrossRef]
  92. Farhan, S.M.K.; Robinson, J.F.; McIntyre, A.D.; Marrosu, M.G.; Ticca, A.F.; Loddo, S.; Carboni, N.; Brancati, F.; Hegele, R.A. A Novel LIPE Nonsense Mutation Found Using Exome Sequencing in Siblings with Late-Onset Familial Partial Lipodystrophy. Can. J. Cardiol. 2014, 30, 1649–1654. [Google Scholar] [CrossRef] [PubMed]
  93. Francis, G.A.; Li, G.; Casey, R.; Wang, J.; Cao, H.; Leff, T.; Hegele, R.A. Peroxisomal Proliferator Activated Receptor-Gamma Deficiency in a Canadian Kindred with Familial Partial Lipodystrophy Type 3 (FPLD3). BMC Med. Genet. 2006, 7, 3. [Google Scholar] [CrossRef]
  94. Gandotra, S.; Le Dour, C.; Bottomley, W.; Cervera, P.; Giral, P.; Reznik, Y.; Charpentier, G.; Auclair, M.; Delépine, M.; Barroso, I.; et al. Perilipin Deficiency and Autosomal Dominant Partial Lipodystrophy. N. Engl. J. Med. 2011, 364, 740–748. [Google Scholar] [CrossRef] [PubMed]
  95. Garg, A.; Kircher, M.; Del Campo, M.; Amato, R.S.; Agarwal, A.K.; University of Washington Center for Mendelian Genomics. Whole Exome Sequencing Identifies de Novo Heterozygous CAV1 Mutations Associated with a Novel Neonatal Onset Lipodystrophy Syndrome. Am. J. Med. Genet. A 2015, 167, 1796–1806. [Google Scholar] [CrossRef]
  96. Guidorizzi, N.R.; Valerio, C.M.; Viola, L.F.; Veras, V.R.; Fernandes, V.O.; da Cruz Paiva Lima, G.E.; Flor, A.C.; Araújo, J.S.; Gonçalves Muniz, R.B.; Moreira, R.O.; et al. Comprehensive Analysis of Morbidity and Mortality Patterns in Familial Partial Lipodystrophy Patients: Insights from a Population Study. Front. Endocrinol. 2024, 15, 1359211. [Google Scholar] [CrossRef]
  97. Guillín-Amarelle, C.; Sánchez-Iglesias, S.; Castro-Pais, A.; Rodriguez-Cañete, L.; Ordóñez-Mayán, L.; Pazos, M.; González-Méndez, B.; Rodríguez-García, S.; Casanueva, F.F.; Fernández-Marmiesse, A.; et al. Type 1 Familial Partial Lipodystrophy: Understanding the Köbberling Syndrome. Endocrine 2016, 54, 411–421. [Google Scholar] [CrossRef] [PubMed]
  98. Hegele, R.A.; Ur, E.; Ransom, T.P.; Cao, H. A Frameshift Mutation in Peroxisome-Proliferator-Activated Receptor-Gamma in Familial Partial Lipodystrophy Subtype 3 (FPLD3; MIM 604367). Clin. Genet. 2006, 70, 360–362. [Google Scholar] [CrossRef]
  99. Herbst, K.L.; Tannock, L.R.; Deeb, S.S.; Purnell, J.Q.; Brunzell, J.D.; Chait, A. Köbberling Type of Familial Partial Lipodystrophy: An Underrecognized Syndrome. Diabetes Care 2003, 26, 1819–1824. [Google Scholar] [CrossRef]
  100. Jéru, I.; Vantyghem, M.-C.; Bismuth, E.; Cervera, P.; Barraud, S.; PLIN1-Study Group; Auclair, M.; Vatier, C.; Lascols, O.; Savage, D.B.; et al. Diagnostic Challenge in PLIN1-Associated Familial Partial Lipodystrophy. J. Clin. Endocrinol. Metab. 2019, 104, 6025–6032. [Google Scholar] [CrossRef]
  101. Kozusko, K.; Tsang, V.; Bottomley, W.; Cho, Y.H.; Gandotra, S.; Mimmack, M.L.; Lim, K.; Isaac, I.; Patel, S.; Saudek, V.; et al. Clinical and Molecular Characterization of a Novel PLIN1 Frameshift Mutation Identified in Patients with Familial Partial Lipodystrophy. Diabetes 2015, 64, 299–310. [Google Scholar] [CrossRef]
  102. Laver, T.W.; Patel, K.A.; Colclough, K.; Curran, J.; Dale, J.; Davis, N.; Savage, D.B.; Flanagan, S.E.; Ellard, S.; Hattersley, A.T.; et al. PLIN1 Haploinsufficiency Is Not Associated With Lipodystrophy. J. Clin. Endocrinol. Metab. 2018, 103, 3225–3230. [Google Scholar] [CrossRef] [PubMed]
  103. Payne, F.; Lim, K.; Girousse, A.; Brown, R.J.; Kory, N.; Robbins, A.; Xue, Y.; Sleigh, A.; Cochran, E.; Adams, C.; et al. Mutations Disrupting the Kennedy Phosphatidylcholine Pathway in Humans with Congenital Lipodystrophy and Fatty Liver Disease. Proc Natl. Acad. Sci. USA 2014, 111, 8901–8906. [Google Scholar] [CrossRef] [PubMed]
  104. Rubio-Cabezas, O.; Puri, V.; Murano, I.; Saudek, V.; Semple, R.K.; Dash, S.; Hyden, C.S.S.; Bottomley, W.; Vigouroux, C.; Magré, J.; et al. Partial Lipodystrophy and Insulin Resistant Diabetes in a Patient with a Homozygous Nonsense Mutation in CIDEC. EMBO Mol. Med. 2009, 1, 280–287. [Google Scholar] [CrossRef]
  105. Schuermans, N.; El Chehadeh, S.; Hemelsoet, D.; Gautheron, J.; Vantyghem, M.-C.; Nouioua, S.; Tazir, M.; Vigouroux, C.; Auclair, M.; Bogaert, E.; et al. Loss of Phospholipase PLAAT3 Causes a Mixed Lipodystrophic and Neurological Syndrome Due to Impaired PPARγ Signaling. Nat. Genet. 2023, 55, 1929–1940. [Google Scholar] [CrossRef] [PubMed]
  106. Semple, R.K.; Chatterjee, V.K.K.; O’Rahilly, S. PPAR Gamma and Human Metabolic Disease. J. Clin. Investig. 2006, 116, 581–589. [Google Scholar] [CrossRef]
  107. Shearin, A.L.; Monks, B.R.; Seale, P.; Birnbaum, M.J. Lack of AKT in Adipocytes Causes Severe Lipodystrophy. Mol. Metab. 2016, 5, 472–479. [Google Scholar] [CrossRef]
  108. Sollier, C.; Capel, E.; Aguilhon, C.; Smirnov, V.; Auclair, M.; Douillard, C.; Ladsous, M.; Defoort-Dhellemmes, S.; Gorwood, J.; Braud, L.; et al. LIPE-Related Lipodystrophic Syndrome: Clinical Features and Disease Modeling Using Adipose Stem Cells. Eur. J. Endocrinol. 2021, 184, 155–168. [Google Scholar] [CrossRef]
  109. Garg, A.; Sankella, S.; Xing, C.; Agarwal, A.K. Whole-Exome Sequencing Identifies ADRA2A Mutation in Atypical Familial Partial Lipodystrophy. JCI Insight 2016, 1, e86870. [Google Scholar] [CrossRef]
  110. Zolotov, S.; Xing, C.; Mahamid, R.; Shalata, A.; Sheikh-Ahmad, M.; Garg, A. Homozygous LIPE Mutation in Siblings with Multiple Symmetric Lipomatosis, Partial Lipodystrophy, and Myopathy. Am. J. Med. Genet. A 2017, 173, 190–194. [Google Scholar] [CrossRef]
  111. Mosbah, H.; Vantyghem, M.-C.; Nobécourt, E.; Andreelli, F.; Archambeaud, F.; Bismuth, E.; Briet, C.; Cartigny, M.; Chevalier, B.; Donadille, B.; et al. Therapeutic Indications and Metabolic Effects of Metreleptin in Patients with Lipodystrophy Syndromes: Real-Life Experience from a National Reference Network. Diabetes Obes. Metab. 2022, 24, 1565–1577. [Google Scholar] [CrossRef] [PubMed]
  112. Park, J.Y.; Javor, E.D.; Cochran, E.K.; DePaoli, A.M.; Gorden, P. Long-Term Efficacy of Leptin Replacement in Patients with Dunnigan-Type Familial Partial Lipodystrophy. Metabolism 2007, 56, 508–516. [Google Scholar] [CrossRef]
  113. Simha, V.; Subramanyam, L.; Szczepaniak, L.; Quittner, C.; Adams-Huet, B.; Snell, P.; Garg, A. Comparison of Efficacy and Safety of Leptin Replacement Therapy in Moderately and Severely Hypoleptinemic Patients with Familial Partial Lipodystrophy of the Dunnigan Variety. J. Clin. Endocrinol. Metab. 2012, 97, 785–792. [Google Scholar] [CrossRef] [PubMed]
  114. Javor, E.D.; Ghany, M.G.; Cochran, E.K.; Oral, E.A.; DePaoli, A.M.; Premkumar, A.; Kleiner, D.E.; Gorden, P. Leptin Reverses Nonalcoholic Steatohepatitis in Patients with Severe Lipodystrophy. Hepatology 2005, 41, 753–760. [Google Scholar] [CrossRef]
  115. Oral, E.A.; Garg, A.; Tami, J.; Huang, E.A.; O’Dea, L.S.L.; Schmidt, H.; Tiulpakov, A.; Mertens, A.; Alexander, V.J.; Watts, L.; et al. Assessment of Efficacy and Safety of Volanesorsen for Treatment of Metabolic Complications in Patients with Familial Partial Lipodystrophy: Results of the BROADEN Study: Volanesorsen in FPLD; The BROADEN Study. J. Clin. Lipidol. 2022, 16, 833–849. [Google Scholar] [CrossRef]
  116. Akinci, B.; Swaidan, M.; Foss-Freitas, M.C.; Luo, Y.; Neidert, A.H.; Hench, R.P.; Chenevert, T.L.; Longcore, A.; Bakker-Arkema, R.; Bisgaier, C.L.; et al. 2214-PUB: An Open-Label Study of Gemcabene in Adults with Familial Partial Lipodystrophy. Diabetes 2020, 69 (Suppl. S1), 2214-PUB. [Google Scholar] [CrossRef]
  117. Foss-Freitas, M.C.; Akinci, B.; Neidert, A.; Bartlett, V.J.; Hurh, E.; Karwatowska-Prokopczuk, E.; Oral, E.A. Selective Targeting of Angiopoietin-like 3 (ANGPTL3) with Vupanorsen for the Treatment of Patients with Familial Partial Lipodystrophy (FPLD): Results of a Proof-of-Concept Study. Lipids Health Dis. 2021, 20, 174. [Google Scholar] [CrossRef]
  118. De Sandre-Giovannoli, A.; Bernard, R.; Cau, P.; Navarro, C.; Amiel, J.; Boccaccio, I.; Lyonnet, S.; Stewart, C.L.; Munnich, A.; Le Merrer, M.; et al. Lamin a Truncation in Hutchinson-Gilford Progeria. Science 2003, 300, 2055. [Google Scholar] [CrossRef] [PubMed]
  119. Eriksson, M.; Brown, W.T.; Gordon, L.B.; Glynn, M.W.; Singer, J.; Scott, L.; Erdos, M.R.; Robbins, C.M.; Moses, T.Y.; Berglund, P.; et al. Recurrent de Novo Point Mutations in Lamin A Cause Hutchinson-Gilford Progeria Syndrome. Nature 2003, 423, 293–298. [Google Scholar] [CrossRef]
  120. Hennekam, R.C.M. Hutchinson-Gilford Progeria Syndrome: Review of the Phenotype. Am. J. Med. Genet. A 2006, 140, 2603–2624. [Google Scholar] [CrossRef]
  121. Mazereeuw-Hautier, J.; Wilson, L.C.; Mohammed, S.; Smallwood, D.; Shackleton, S.; Atherton, D.J.; Harper, J.I. Hutchinson-Gilford Progeria Syndrome: Clinical Findings in Three Patients Carrying the G608G Mutation in LMNA and Review of the Literature. Br. J. Dermatol. 2007, 156, 1308–1314. [Google Scholar] [CrossRef] [PubMed]
  122. Merideth, M.A.; Gordon, L.B.; Clauss, S.; Sachdev, V.; Smith, A.C.M.; Perry, M.B.; Brewer, C.C.; Zalewski, C.; Kim, H.J.; Solomon, B.; et al. Phenotype and Course of Hutchinson-Gilford Progeria Syndrome. N. Engl. J. Med. 2008, 358, 592–604. [Google Scholar] [CrossRef] [PubMed]
  123. Gordon, L.B.; Brown, W.T.; Collins, F.S. Hutchinson-Gilford Progeria Syndrome. 2003 Dec 12 [Updated 2023 Oct 19]; Adam, M.P., Feldman, J., Mirzaa, G.M., Pagon, R.A., Wallace, S.E., Bean, L.G.H., Gripp, K.W., Amemiya, A., Eds.; GeneReviews® [Internet] 1993–2024; University of Washington: Seattle, WA, USA, 2003. [Google Scholar]
  124. Stehbens, W.E.; Wakefield, S.J.; Gilbert-Barness, E.; Olson, R.E.; Ackerman, J. Histological and Ultrastructural Features of Atherosclerosis in Progeria. Cardiovasc. Pathol. 1999, 8, 29–39. [Google Scholar] [CrossRef]
  125. Gordon, L.B.; Harten, I.A.; Patti, M.E.; Lichtenstein, A.H. Reduced Adiponectin and HDL Cholesterol without Elevated C-Reactive Protein: Clues to the Biology of Premature Atherosclerosis in Hutchinson-Gilford Progeria Syndrome. J. Pediatr. 2005, 146, 336–341. [Google Scholar] [CrossRef]
  126. Olive, M.; Harten, I.; Mitchell, R.; Beers, J.K.; Djabali, K.; Cao, K.; Erdos, M.R.; Blair, C.; Funke, B.; Smoot, L.; et al. Cardiovascular Pathology in Hutchinson-Gilford Progeria: Correlation with the Vascular Pathology of Aging. Arterioscler. Thromb. Vasc. Biol. 2010, 30, 2301–2309. [Google Scholar] [CrossRef]
  127. Gerhard-Herman, M.; Smoot, L.B.; Wake, N.; Kieran, M.W.; Kleinman, M.E.; Miller, D.T.; Schwartzman, A.; Giobbie-Hurder, A.; Neuberg, D.; Gordon, L.B. Mechanisms of Premature Vascular Aging in Children with Hutchinson-Gilford Progeria Syndrome. Hypertension 2012, 59, 92–97. [Google Scholar] [CrossRef]
  128. Osorio, F.G.; Navarro, C.L.; Cadiñanos, J.; López-Mejía, I.C.; Quirós, P.M.; Bartoli, C.; Rivera, J.; Tazi, J.; Guzmán, G.; Varela, I.; et al. Splicing-Directed Therapy in a New Mouse Model of Human Accelerated Aging. Sci. Transl. Med. 2011, 3, 106ra107. [Google Scholar] [CrossRef] [PubMed]
  129. Varga, R.; Eriksson, M.; Erdos, M.R.; Olive, M.; Harten, I.; Kolodgie, F.; Capell, B.C.; Cheng, J.; Faddah, D.; Perkins, S.; et al. Progressive Vascular Smooth Muscle Cell Defects in a Mouse Model of Hutchinson-Gilford Progeria Syndrome. Proc. Natl. Acad. Sci. USA 2006, 103, 3250–3255. [Google Scholar] [CrossRef]
  130. Villa-Bellosta, R.; Rivera-Torres, J.; Osorio, F.G.; Acín-Pérez, R.; Enriquez, J.A.; López-Otín, C.; Andrés, V. Defective Extracellular Pyrophosphate Metabolism Promotes Vascular Calcification in a Mouse Model of Hutchinson-Gilford Progeria Syndrome That Is Ameliorated on Pyrophosphate Treatment. Circulation 2013, 127, 2442–2451. [Google Scholar] [CrossRef]
  131. Fong, L.G.; Frost, D.; Meta, M.; Qiao, X.; Yang, S.H.; Coffinier, C.; Young, S.G. A Protein Farnesyltransferase Inhibitor Ameliorates Disease in a Mouse Model of Progeria. Science 2006, 311, 1621–1623. [Google Scholar] [CrossRef]
  132. Gordon, L.B.; Kleinman, M.E.; Miller, D.T.; Neuberg, D.S.; Giobbie-Hurder, A.; Gerhard-Herman, M.; Smoot, L.B.; Gordon, C.M.; Cleveland, R.; Snyder, B.D.; et al. Clinical Trial of a Farnesyltransferase Inhibitor in Children with Hutchinson-Gilford Progeria Syndrome. Proc. Natl. Acad. Sci. USA 2012, 109, 16666–16671. [Google Scholar] [CrossRef]
  133. Yang, S.H.; Meta, M.; Qiao, X.; Frost, D.; Bauch, J.; Coffinier, C.; Majumdar, S.; Bergo, M.O.; Young, S.G.; Fong, L.G.A. Farnesyltransferase Inhibitor Improves Disease Phenotypes in Mice with a Hutchinson-Gilford Progeria Syndrome Mutation. J. Clin. Investig. 2006, 116, 2115–2121. [Google Scholar] [CrossRef]
  134. Gordon, L.B.; Shappell, H.; Massaro, J.; D’Agostino, R.B.; Brazier, J.; Campbell, S.E.; Kleinman, M.E.; Kieran, M.W. Association of Lonafarnib Treatment vs No Treatment With Mortality Rate in Patients With Hutchinson-Gilford Progeria Syndrome. JAMA 2018, 319, 1687–1695. [Google Scholar] [CrossRef] [PubMed]
  135. Gordon, L.B.; Massaro, J.; D’Agostino, R.B.; Campbell, S.E.; Brazier, J.; Brown, W.T.; Kleinman, M.E.; Kieran, M.W.; Progeria Clinical Trials Collaborative. Impact of Farnesylation Inhibitors on Survival in Hutchinson-Gilford Progeria Syndrome. Circulation 2014, 130, 27–34. [Google Scholar] [CrossRef]
  136. Erdos, M.R.; Cabral, W.A.; Tavarez, U.L.; Cao, K.; Gvozdenovic-Jeremic, J.; Narisu, N.; Zerfas, P.M.; Crumley, S.; Boku, Y.; Hanson, G.; et al. A Targeted Antisense Therapeutic Approach for Hutchinson-Gilford Progeria Syndrome. Nat. Med. 2021, 27, 536–545. [Google Scholar] [CrossRef]
  137. Santiago-Fernández, O.; Osorio, F.G.; Quesada, V.; Rodríguez, F.; Basso, S.; Maeso, D.; Rolas, L.; Barkaway, A.; Nourshargh, S.; Folgueras, A.R.; et al. Development of a CRISPR/Cas9-Based Therapy for Hutchinson-Gilford Progeria Syndrome. Nat. Med. 2019, 25, 423–426. [Google Scholar] [CrossRef] [PubMed]
  138. Doubaj, Y.; De Sandre-Giovannoli, A.; Vera, E.-V.; Navarro, C.L.; Elalaoui, S.C.; Tajir, M.; Lévy, N.; Sefiani, A. An Inherited LMNA Gene Mutation in Atypical Progeria Syndrome. Am. J. Med. Genet. A 2012, 158, 2881–2887. [Google Scholar] [CrossRef] [PubMed]
  139. Garg, A.; Subramanyam, L.; Agarwal, A.K.; Simha, V.; Levine, B.; D’Apice, M.R.; Novelli, G.; Crow, Y. Atypical Progeroid Syndrome Due to Heterozygous Missense LMNA Mutations. J. Clin. Endocrinol. Metab. 2009, 94, 4971–4983. [Google Scholar] [CrossRef] [PubMed]
  140. Hussain, I.; Patni, N.; Ueda, M.; Sorkina, E.; Valerio, C.M.; Cochran, E.; Brown, R.J.; Peeden, J.; Tikhonovich, Y.; Tiulpakov, A.; et al. A Novel Generalized Lipodystrophy-Associated Progeroid Syndrome Due to Recurrent Heterozygous LMNA p.T10I Mutation. J. Clin. Endocrinol. Metab. 2018, 103, 1005–1014. [Google Scholar] [CrossRef]
  141. Hussain, I.; Jin, R.R.; Baum, H.B.A.; Greenfield, J.R.; Devery, S.; Xing, C.; Hegele, R.A.; Carranza-Leon, B.G.; Linton, M.F.; Vuitch, F.; et al. Multisystem Progeroid Syndrome With Lipodystrophy, Cardiomyopathy, and Nephropathy Due to an LMNA p.R349W Variant. J. Endocr. Soc. 2020, 4, bvaa104. [Google Scholar] [CrossRef]
  142. Magno, S.; Ceccarini, G.; Pelosini, C.; Ferrari, F.; Prodam, F.; Gilio, D.; Maffei, M.; Sessa, M.R.; Barison, A.; Ciccarone, A.; et al. Atypical Progeroid Syndrome and Partial Lipodystrophy Due to LMNA Gene p.R349W Mutation. J. Endocr. Soc. 2020, 4, bvaa108. [Google Scholar] [CrossRef]
  143. Motegi, S.; Yokoyama, Y.; Uchiyama, A.; Ogino, S.; Takeuchi, Y.; Yamada, K.; Hattori, T.; Hashizume, H.; Ishikawa, Y.; Goto, M.; et al. First Japanese Case of Atypical Progeroid Syndrome/Atypical Werner Syndrome with Heterozygous LMNA Mutation. J. Dermatol. 2014, 41, 1047–1052. [Google Scholar] [CrossRef] [PubMed]
  144. Csoka, A.B.; Cao, H.; Sammak, P.J.; Constantinescu, D.; Schatten, G.P.; Hegele, R.A. Novel Lamin A/C Gene (LMNA) Mutations in Atypical Progeroid Syndromes. J. Med. Genet. 2004, 41, 304–308. [Google Scholar] [CrossRef] [PubMed]
  145. Mory, P.B.; Crispim, F.; Kasamatsu, T.; Gabbay, M.A.L.; Dib, S.A.; Moisés, R.S. Atypical Generalized Lipoatrophy and Severe Insulin Resistance Due to a Heterozygous LMNA p.T10I Mutation. Arq. Bras. Endocrinol. Metabol. 2008, 52, 1252–1256. [Google Scholar] [CrossRef] [PubMed]
  146. Saadi, A.; Navarro, C.; Ozalp, O.; Lourenco, C.M.; Fayek, R.; Da Silva, N.; Chaouch, A.; Benahmed, M.; Kubisch, C.; Munnich, A.; et al. A Recurrent Homozygous LMNA Missense Variant p.Thr528Met Causes Atypical Progeroid Syndrome Characterized by Mandibuloacral Dysostosis, Severe Muscular Dystrophy, and Skeletal Deformities. Am. J. Med. Genet. A 2023, 191, 2274–2289. [Google Scholar] [CrossRef]
  147. Araujo-Vilar, D.; Sánchez-Iglesias, S.; Guillín-Amarelle, C.; Castro, A.; Lage, M.; Pazos, M.; Rial, J.M.; Blasco, J.; Guillén-Navarro, E.; Domingo-Jiménez, R.; et al. Recombinant Human Leptin Treatment in Genetic Lipodystrophic Syndromes: The Long-Term Spanish Experience. Endocrine 2015, 49, 139–147. [Google Scholar] [CrossRef] [PubMed]
  148. Soyaltin, U.E.; Simsir, I.Y.; Akinci, B.; Altay, C.; Adiyaman, S.C.; Lee, K.; Onay, H.; Oral, E.A. Homozygous LMNA p.R582H Pathogenic Variant Reveals Increasing Effect on the Severity of Fat Loss in Lipodystrophy. Clin. Diabetes. Endocrinol. 2020, 6, 13. [Google Scholar] [CrossRef]
  149. Ng, K.K.; Kaye, G. A Case of Lamin C Gene-Mutation with Preserved Systolic Function and Ventricular Dysrrhythmia. Australas. Med. J. 2013, 6, 75–78. [Google Scholar] [CrossRef]
  150. Benedetti, S.; Bertini, E.; Iannaccone, S.; Angelini, C.; Trisciani, M.; Toniolo, D.; Sferrazza, B.; Carrera, P.; Comi, G.; Ferrari, M.; et al. Dominant LMNA Mutations Can Cause Combined Muscular Dystrophy and Peripheral Neuropathy. J. Neurol. Neurosurg. Psychiatry 2005, 76, 1019–1021. [Google Scholar] [CrossRef]
  151. Patni, N.; Xing, C.; Agarwal, A.K.; Garg, A. Juvenile-Onset Generalized Lipodystrophy Due to a Novel Heterozygous Missense LMNA Mutation Affecting Lamin C. Am. J. Med. Genet. A 2017, 173, 2517–2521. [Google Scholar] [CrossRef]
  152. Kane, M.S.; Lindsay, M.E.; Judge, D.P.; Barrowman, J.; Ap Rhys, C.; Simonson, L.; Dietz, H.C.; Michaelis, S. LMNA-Associated Cardiocutaneous Progeria: An Inherited Autosomal Dominant Premature Aging Syndrome with Late Onset. Am. J. Med. Genet. A 2013, 161, 1599–1611. [Google Scholar] [CrossRef]
  153. Cenni, V.; D’Apice, M.R.; Garagnani, P.; Columbaro, M.; Novelli, G.; Franceschi, C.; Lattanzi, G. Mandibuloacral Dysplasia: A Premature Ageing Disease with Aspects of Physiological Ageing. Ageing Res. Rev. 2018, 42, 1–13. [Google Scholar] [CrossRef]
  154. Novelli, G.; Muchir, A.; Sangiuolo, F.; Helbling-Leclerc, A.; D’Apice, M.R.; Massart, C.; Capon, F.; Sbraccia, P.; Federici, M.; Lauro, R.; et al. Mandibuloacral Dysplasia Is Caused by a Mutation in LMNA-Encoding Lamin A/C. Am. J. Hum. Genet. 2002, 71, 426–431. [Google Scholar] [CrossRef] [PubMed]
  155. Lombardi, F.; Gullotta, F.; Columbaro, M.; Filareto, A.; D’Adamo, M.; Vielle, A.; Guglielmi, V.; Nardone, A.M.; Azzolini, V.; Grosso, E.; et al. Compound Heterozygosity for Mutations in LMNA in a Patient with a Myopathic and Lipodystrophic Mandibuloacral Dysplasia Type A Phenotype. J. Clin. Endocrinol. Metab. 2007, 92, 4467–4471. [Google Scholar] [CrossRef]
  156. Simha, V.; Agarwal, A.K.; Oral, E.A.; Fryns, J.-P.; Garg, A. Genetic and Phenotypic Heterogeneity in Patients with Mandibuloacral Dysplasia-Associated Lipodystrophy. J. Clin. Endocrinol. Metab. 2003, 88, 2821–2824. [Google Scholar] [CrossRef] [PubMed]
  157. Agarwal, A.K.; Fryns, J.-P.; Auchus, R.J.; Garg, A. Zinc Metalloproteinase, ZMPSTE24, Is Mutated in Mandibuloacral Dysplasia. Hum. Mol. Genet. 2003, 12, 1995–2001. [Google Scholar] [CrossRef]
  158. Miyoshi, Y.; Akagi, M.; Agarwal, A.K.; Namba, N.; Kato-Nishimura, K.; Mohri, I.; Yamagata, M.; Nakajima, S.; Mushiake, S.; Shima, M.; et al. Severe Mandibuloacral Dysplasia Caused by Novel Compound Heterozygous ZMPSTE24 Mutations in Two Japanese Siblings. Clin. Genet. 2008, 73, 535–544. [Google Scholar] [CrossRef]
  159. Young, L.W.; Radebaugh, J.F.; Rubin, P.; Sensenbrenner, J.A.; Fiorelli, G.; McKusick, V.A. New Syndrome Manifested by Mandibular Hypoplasia, Acroosteolysis, Stiff Joints and Cutaneous Atrophy (Mandibuloacral Dysplasia) in Two Unrelated Boys. Birth Defects Orig. Artic. Ser. 1971, 7, 291–297. [Google Scholar]
  160. Cunningham, V.J.; D’Apice, M.R.; Licata, N.; Novelli, G.; Cundy, T. Skeletal Phenotype of Mandibuloacral Dysplasia Associated with Mutations in ZMPSTE24. Bone 2010, 47, 591–597. [Google Scholar] [CrossRef]
  161. Camozzi, D.; D’Apice, M.R.; Schena, E.; Cenni, V.; Columbaro, M.; Capanni, C.; Maraldi, N.M.; Squarzoni, S.; Ortolani, M.; Novelli, G.; et al. Altered Chromatin Organization and SUN2 Localization in Mandibuloacral Dysplasia Are Rescued by Drug Treatment. Histochem. Cell. Biol. 2012, 138, 643–651. [Google Scholar] [CrossRef] [PubMed]
  162. Lombardi, F.; Fasciglione, G.F.; D’Apice, M.R.; Vielle, A.; D’Adamo, M.; Sbraccia, P.; Marini, S.; Borgiani, P.; Coletta, M.; Novelli, G. Increased Release and Activity of Matrix Metalloproteinase-9 in Patients with Mandibuloacral Dysplasia Type A, a Rare Premature Ageing Syndrome. Clin. Genet. 2008, 74, 374–383. [Google Scholar] [CrossRef]
  163. Jéru, I.; Nabil, A.; El-Makkawy, G.; Lascols, O.; Vigouroux, C.; Abdalla, E. Two Decades after Mandibuloacral Dysplasia Discovery: Additional Cases and Comprehensive View of Disease Characteristics. Genes 2021, 12, 1508. [Google Scholar] [CrossRef] [PubMed]
  164. Ozer, L.; Unsal, E.; Aktuna, S.; Baltaci, V.; Celikkol, P.; Akyigit, F.; Sen, A.; Ayvaz, O.; Balci, S. Mandibuloacral Dysplasia and LMNA A529V Mutation in Turkish Patients with Severe Skeletal Changes and Absent Breast Development. Clin. Dysmorphol. 2016, 25, 91–97. [Google Scholar] [CrossRef]
  165. Avnet, S.; Pallotta, R.; Perut, F.; Baldini, N.; Pittis, M.G.; Saponari, A.; Lucarelli, E.; Dozza, B.; Greggi, T.; Maraldi, N.M.; et al. Osteoblasts from a Mandibuloacral Dysplasia Patient Induce Human Blood Precursors to Differentiate into Active Osteoclasts. Biochim. Biophys. Acta 2011, 1812, 711–718. [Google Scholar] [CrossRef]
  166. Garavelli, L.; D’Apice, M.R.; Rivieri, F.; Bertoli, M.; Wischmeijer, A.; Gelmini, C.; De Nigris, V.; Albertini, E.; Rosato, S.; Virdis, R.; et al. Mandibuloacral Dysplasia Type A in Childhood. Am. J. Med. Genet. A 2009, 149, 2258–2264. [Google Scholar] [CrossRef] [PubMed]
  167. Camozzi, D.; Capanni, C.; Cenni, V.; Mattioli, E.; Columbaro, M.; Squarzoni, S.; Lattanzi, G. Diverse Lamin-Dependent Mechanisms Interact to Control Chromatin Dynamics. Focus on Laminopathies. Nucleus 2014, 5, 427–440. [Google Scholar] [CrossRef] [PubMed]
  168. Evangelisti, C.; Bernasconi, P.; Cavalcante, P.; Cappelletti, C.; D’Apice, M.R.; Sbraccia, P.; Novelli, G.; Prencipe, S.; Lemma, S.; Baldini, N.; et al. Modulation of TGFbeta 2 Levels by Lamin A in U2-OS Osteoblast-like Cells: Understanding the Osteolytic Process Triggered by Altered Lamins. Oncotarget 2015, 6, 7424–7437. [Google Scholar] [CrossRef]
  169. Fong, L.G.; Ng, J.K.; Meta, M.; Coté, N.; Yang, S.H.; Stewart, C.L.; Sullivan, T.; Burghardt, A.; Majumdar, S.; Reue, K.; et al. Heterozygosity for Lmna Deficiency Eliminates the Progeria-like Phenotypes in Zmpste24-Deficient Mice. Proc. Natl. Acad. Sci. USA 2004, 101, 18111–18116. [Google Scholar] [CrossRef]
  170. Navarro, C.L.; De Sandre-Giovannoli, A.; Bernard, R.; Boccaccio, I.; Boyer, A.; Geneviève, D.; Hadj-Rabia, S.; Gaudy-Marqueste, C.; Smitt, H.S.; Vabres, P.; et al. Lamin A and ZMPSTE24 (FACE-1) Defects Cause Nuclear Disorganization and Identify Restrictive Dermopathy as a Lethal Neonatal Laminopathy. Hum. Mol. Genet. 2004, 13, 2493–2503. [Google Scholar] [CrossRef]
  171. Simha, V.; Garg, A. Body Fat Distribution and Metabolic Derangements in Patients with Familial Partial Lipodystrophy Associated with Mandibuloacral Dysplasia. J. Clin. Endocrinol. Metab. 2002, 87, 776–785. [Google Scholar] [CrossRef]
  172. Hitzert, M.M.; van der Crabben, S.N.; Baldewsingh, G.; van Amstel, H.K.P.; van den Wijngaard, A.; van Ravenswaaij-Arts, C.M.A.; Zijlmans, C.W.R. Mandibuloacral Dysplasia Type B (MADB): A Cohort of Eight Patients from Suriname with a Homozygous Founder Mutation in ZMPSTE24 (FACE1), Clinical Diagnostic Criteria and Management Guidelines. Orphanet J. Rare Dis. 2019, 14, 294. [Google Scholar] [CrossRef] [PubMed]
  173. Haye, D.; Dridi, H.; Levy, J.; Lambert, V.; Lambert, M.; Agha, M.; Adjimi, F.; Kohlhase, J.; Lipsker, D.; Verloes, A. Failure of Ossification of the Occipital Bone in Mandibuloacral Dysplasia Type B. Am. J. Med. Genet. A 2016, 170, 2750–2755. [Google Scholar] [CrossRef] [PubMed]
  174. Ahmad, Z.; Zackai, E.; Medne, L.; Garg, A. Early Onset Mandibuloacral Dysplasia Due to Compound Heterozygous Mutations in ZMPSTE24. Am. J. Med. Genet. A 2010, 152, 2703–2710. [Google Scholar] [CrossRef]
  175. Varela, I.; Pereira, S.; Ugalde, A.P.; Navarro, C.L.; Suárez, M.F.; Cau, P.; Cadiñanos, J.; Osorio, F.G.; Foray, N.; Cobo, J.; et al. Combined Treatment with Statins and Aminobisphosphonates Extends Longevity in a Mouse Model of Human Premature Aging. Nat. Med. 2008, 14, 767–772. [Google Scholar] [CrossRef]
  176. Hartinger, R.; Lederer, E.-M.; Schena, E.; Lattanzi, G.; Djabali, K. Impact of Combined Baricitinib and FTI Treatment on Adipogenesis in Hutchinson-Gilford Progeria Syndrome and Other Lipodystrophic Laminopathies. Cells 2023, 12, 1350. [Google Scholar] [CrossRef] [PubMed]
  177. Akinci, B.; Sankella, S.; Gilpin, C.; Ozono, K.; Garg, A.; Agarwal, A.K. Progeroid Syndrome Patients with ZMPSTE24 Deficiency Could Benefit When Treated with Rapamycin and Dimethylsulfoxide. Cold Spring Harb. Mol. Case Stud. 2017, 3, a001339. [Google Scholar] [CrossRef]
  178. Odinammadu, K.O.; Shilagardi, K.; Tuminelli, K.; Judge, D.P.; Gordon, L.B.; Michaelis, S. The Farnesyl Transferase Inhibitor (FTI) Lonafarnib Improves Nuclear Morphology in ZMPSTE24-Deficient Fibroblasts from Patients with the Progeroid Disorder MAD-B. Nucleus 2023, 14, 2288476. [Google Scholar] [CrossRef] [PubMed]
  179. Harhouri, K.; Navarro, C.; Baquerre, C.; Da Silva, N.; Bartoli, C.; Casey, F.; Mawuse, G.K.; Doubaj, Y.; Lévy, N.; De Sandre-Giovannoli, A. Antisense-Based Progerin Downregulation in HGPS-Like Patients’ Cells. Cells 2016, 5, 31. [Google Scholar] [CrossRef]
  180. Fisher, H.G.; Patni, N.; Scheuerle, A.E. An Additional Case of Néstor-Guillermo Progeria Syndrome Diagnosed in Early Childhood. Am. J. Med. Genet. A 2020, 182, 2399–2402. [Google Scholar] [CrossRef]
  181. Puente, X.S.; Quesada, V.; Osorio, F.G.; Cabanillas, R.; Cadiñanos, J.; Fraile, J.M.; Ordóñez, G.R.; Puente, D.A.; Gutiérrez-Fernández, A.; Fanjul-Fernández, M.; et al. Exome Sequencing and Functional Analysis Identifies BANF1 Mutation as the Cause of a Hereditary Progeroid Syndrome. Am. J. Hum. Genet. 2011, 88, 650–656. [Google Scholar] [CrossRef]
  182. Marcelot, A.; Rodriguez-Tirado, F.; Cuniasse, P.; Joiner, M.-L.; Miron, S.; Soshnev, A.A.; Fang, M.; Pufall, M.A.; Mathews, K.D.; Moore, S.A.; et al. A De Novo Sequence Variant in Barrier-to-Autointegration Factor Is Associated with Dominant Motor Neuronopathy. Cells 2023, 12, 847. [Google Scholar] [CrossRef] [PubMed]
  183. Guillín-Amarelle, C.; Sánchez-Iglesias, S.; Araújo-Vilar, D. Uncommon lipodystrophic syndromes. Med. Clin. 2015, 144, 80–87. [Google Scholar] [CrossRef] [PubMed]
  184. Hennekam, R.C.M. Pathophysiology of Premature Aging Characteristics in Mendelian Progeroid Disorders. Eur. J. Med. Genet. 2020, 63, 104028. [Google Scholar] [CrossRef] [PubMed]
  185. Neer, R.M.; Arnaud, C.D.; Zanchetta, J.R.; Prince, R.; Gaich, G.A.; Reginster, J.Y.; Hodsman, A.B.; Eriksen, E.F.; Ish-Shalom, S.; Genant, H.K.; et al. Effect of Parathyroid Hormone (1-34) on Fractures and Bone Mineral Density in Postmenopausal Women with Osteoporosis. N. Engl. J. Med. 2001, 344, 1434–1441. [Google Scholar] [CrossRef] [PubMed]
  186. Capell, B.C.; Olive, M.; Erdos, M.R.; Cao, K.; Faddah, D.A.; Tavarez, U.L.; Conneely, K.N.; Qu, X.; San, H.; Ganesh, S.K.; et al. A Farnesyltransferase Inhibitor Prevents Both the Onset and Late Progression of Cardiovascular Disease in a Progeria Mouse Model. Proc. Natl. Acad. Sci. USA 2008, 105, 15902–15907. [Google Scholar] [CrossRef]
  187. Worman, H.J.; Fong, L.G.; Muchir, A.; Young, S.G. Laminopathies and the Long Strange Trip from Basic Cell Biology to Therapy. J. Clin. Investig. 2009, 119, 1825–1836. [Google Scholar] [CrossRef]
  188. López-Otín, C.; Blasco, M.A.; Partridge, L.; Serrano, M.; Kroemer, G. The Hallmarks of Aging. Cell 2013, 153, 1194–1217. [Google Scholar] [CrossRef]
  189. Worman, H.J.; Michaelis, S. Prelamin A and ZMPSTE24 in Premature and Physiological Aging. Nucleus 2023, 14, 2270345. [Google Scholar] [CrossRef]
  190. Akinci, B.; Oral, E.A.; Neidert, A.; Rus, D.; Cheng, W.Y.; Thompson-Leduc, P.; Cheung, H.C.; Bradt, P.; Foss de Freitas, M.C.; Montenegro, R.M.; et al. Comorbidities and Survival in Patients With Lipodystrophy: An International Chart Review Study. J. Clin. Endocrinol. Metab. 2019, 104, 5120–5135. [Google Scholar] [CrossRef]
  191. Tu, Y.; Sánchez-Iglesias, S.; Araújo-Vilar, D.; Fong, L.G.; Young, S.G. LMNA Missense Mutations Causing Familial Partial Lipodystrophy Do Not Lead to an Accumulation of Prelamin A. Nucleus 2016, 7, 512–521. [Google Scholar] [CrossRef]
  192. Carboni, N.; Politano, L.; Floris, M.; Mateddu, A.; Solla, E.; Olla, S.; Maggi, L.; Antonietta Maioli, M.; Piras, R.; Cocco, E.; et al. Overlapping Syndromes in Laminopathies: A Meta-Analysis of the Reported Literature. Acta Myol. 2013, 32, 7–17. [Google Scholar] [PubMed]
  193. Fernández-Pombo, A.; Ossandon-Otero, J.A.; Guillín-Amarelle, C.; Sánchez-Iglesias, S.; Castro, A.I.; González-Méndez, B.; Rodríguez-García, S.; Rodriguez-Cañete, L.; Casanueva, F.F.; Araújo-Vilar, D. Bone Mineral Density in Familial Partial Lipodystrophy. Clin. Endocrinol. 2018, 88, 44–50. [Google Scholar] [CrossRef] [PubMed]
  194. Jackson, S.N.; Howlett, T.A.; McNally, P.G.; O’Rahilly, S.; Trembath, R.C. Dunnigan-Kobberling Syndrome: An Autosomal Dominant Form of Partial Lipodystrophy. QJM 1997, 90, 27–36. [Google Scholar] [CrossRef] [PubMed]
  195. Araújo-Vilar, D.; Fernández-Pombo, A.; Cobelo-Gómez, S.; Castro, A.I.; Sánchez-Iglesias, S. Lipodystrophy-Associated Progeroid Syndromes. Hormones 2022, 21, 555–571. [Google Scholar] [CrossRef]
  196. Kuo, F.-C.; Huang, Y.-C.; Yen, M.-R.; Lee, C.-H.; Hsu, K.-F.; Yang, H.-Y.; Wu, L.-W.; Lu, C.-H.; Hsu, Y.-J.; Chen, P.-Y. Aberrant Overexpression of HOTAIR Inhibits Abdominal Adipogenesis through Remodelling of Genome-Wide DNA Methylation and Transcription. Mol. Metab. 2022, 60, 101473. [Google Scholar] [CrossRef]
  197. Patni, N.; Garg, A. Congenital Generalized Lipodystrophies--New Insights into Metabolic Dysfunction. Nat. Rev. Endocrinol. 2015, 11, 522–534. [Google Scholar] [CrossRef]
  198. Gorden, P.; Lupsa, B.C.; Chong, A.Y.; Lungu, A.O. Is There a Human Model for the “metabolic Syndrome” with a Defined Aetiology? Diabetologia 2010, 53, 1534–1536. [Google Scholar] [CrossRef]
  199. Araújo-Vilar, D.; Santini, F. Diagnosis and Treatment of Lipodystrophy: A Step-by-Step Approach. J. Endocrinol. Investig. 2019, 42, 61–73. [Google Scholar] [CrossRef]
Figure 1. Phenotypic features of patients with lipodystrophic laminopathies. (A) 48-year-old woman with Dunnigan disease and classic phenotype due to the p.(Arg482Trp) variant in exon 8 of the LMNA gene (FPLD2 #1); (B) 62-year-old woman with Dunnigan disease and atypical phenotype due to the p.(Thr528Met) variant in exon 9 of the LMNA gene (FPLD2 #2); (C) Classic FPLD2 is usually characterised by the accumulation of fat in the face and neck, giving a Cushingoid appearance; (D) Signs of insulin resistance such as acanthosis nigricans are common features in FPLD2; (E) The presence of subcutaneous lipomas and a concordant phenotype can guide the diagnosis in patients with FPLD2, as is the case of this subject, who developed lipomas within 3 months of follow-up; (F) 7-year-old boy with Hutchinson-Gilford progeria syndrome diagnosed at birth due to the 1822G > A, p.(G608S) variant in the LMNA gene (HGPS #1), showing the following phenotypical characteristics: generalised lipodystrophy, generalised alopecia, leucomelanodermal macules affecting the entire body, prominent cranial venous tree (H), bulging eyes due to the absence of retroorbital fat, small and sharp nose, small ears of normal implantation, dental malposition and absence of teeth, micrognathia and nasal voice. He presented a distended abdomen, with a reducible umbilical hernia and no hepatosplenomegaly. He showed marked veins in the lower limbs. Apparently normal clavicles, nail dysplasia and coxa valga could also be observed. Regarding comorbidities, he presented recurrent infections, bronchial asthma, subclinical hypothyroidism, myopia and, at the age of 5, he was hospitalised due to an occlusive dissection of the left internal carotid artery; (G) 16-year-old woman with Hutchinson-Gilford progeria syndrome diagnosed at birth due to the c.1824C > T, p.(G608G) variant in the LMNA gene (HGPS #2), showing the following phenotypical characteristics: generalised lipodystrophy, cutaneous sclerosis, generalised alopecia, small and sharp nose, small mouth with dental crowding and micrognathia (I). She presented apparently normal clavicles, dysplastic nails, joint stiffness and coxa valga. Regarding comorbidities, she only showed mild aortic insufficiency. However, at the age of 16 she presented an acute myocardial infarction; (K) 17-year-old man with LMNA-atypical progeroid syndrome diagnosed at 6 months of age due to the heterozygous c.29C > T, p.(Thr10Ile) variant in the LMNA gene, showing the following phenotypic characteristics: generalised lipodystrophy affecting palms and soles (J,L), with thin skin, leucomelanodermal macular lesions (M) and a progeroid facies with proptotic eyeballs as a consequence of the probable absence of retro-orbital fat, a pointed nasal pyramid, high-pitched voice and slight crowding of teeth. Hepatomegaly and splenomegaly were palpable. He had joint contractures affecting the upper and lower limbs and metatarsophalangeal stiffness. He presented pseudodislocation of the ankle joint which conditions talus-valgus feet. There was no resorption of the distal phalanges and he had normal nails. He had mild scoliosis, normal clavicles and no mandibular hypoplasia. He presented phlebomegaly in the limbs (L). Regarding comorbidities, he was diagnosed with diabetes mellitus, hypertriglyceridaemia, dilated cardiomyopathy with moderate pulmonary hypertension, for which he underwent a heart transplant, and he developed central nervous system lymphoma following the immunosuppressive therapy received. The images included here are of patients from our Lipodystrophy Unit (UETeM reference centre) and have not previously been published elsewhere.
Figure 1. Phenotypic features of patients with lipodystrophic laminopathies. (A) 48-year-old woman with Dunnigan disease and classic phenotype due to the p.(Arg482Trp) variant in exon 8 of the LMNA gene (FPLD2 #1); (B) 62-year-old woman with Dunnigan disease and atypical phenotype due to the p.(Thr528Met) variant in exon 9 of the LMNA gene (FPLD2 #2); (C) Classic FPLD2 is usually characterised by the accumulation of fat in the face and neck, giving a Cushingoid appearance; (D) Signs of insulin resistance such as acanthosis nigricans are common features in FPLD2; (E) The presence of subcutaneous lipomas and a concordant phenotype can guide the diagnosis in patients with FPLD2, as is the case of this subject, who developed lipomas within 3 months of follow-up; (F) 7-year-old boy with Hutchinson-Gilford progeria syndrome diagnosed at birth due to the 1822G > A, p.(G608S) variant in the LMNA gene (HGPS #1), showing the following phenotypical characteristics: generalised lipodystrophy, generalised alopecia, leucomelanodermal macules affecting the entire body, prominent cranial venous tree (H), bulging eyes due to the absence of retroorbital fat, small and sharp nose, small ears of normal implantation, dental malposition and absence of teeth, micrognathia and nasal voice. He presented a distended abdomen, with a reducible umbilical hernia and no hepatosplenomegaly. He showed marked veins in the lower limbs. Apparently normal clavicles, nail dysplasia and coxa valga could also be observed. Regarding comorbidities, he presented recurrent infections, bronchial asthma, subclinical hypothyroidism, myopia and, at the age of 5, he was hospitalised due to an occlusive dissection of the left internal carotid artery; (G) 16-year-old woman with Hutchinson-Gilford progeria syndrome diagnosed at birth due to the c.1824C > T, p.(G608G) variant in the LMNA gene (HGPS #2), showing the following phenotypical characteristics: generalised lipodystrophy, cutaneous sclerosis, generalised alopecia, small and sharp nose, small mouth with dental crowding and micrognathia (I). She presented apparently normal clavicles, dysplastic nails, joint stiffness and coxa valga. Regarding comorbidities, she only showed mild aortic insufficiency. However, at the age of 16 she presented an acute myocardial infarction; (K) 17-year-old man with LMNA-atypical progeroid syndrome diagnosed at 6 months of age due to the heterozygous c.29C > T, p.(Thr10Ile) variant in the LMNA gene, showing the following phenotypic characteristics: generalised lipodystrophy affecting palms and soles (J,L), with thin skin, leucomelanodermal macular lesions (M) and a progeroid facies with proptotic eyeballs as a consequence of the probable absence of retro-orbital fat, a pointed nasal pyramid, high-pitched voice and slight crowding of teeth. Hepatomegaly and splenomegaly were palpable. He had joint contractures affecting the upper and lower limbs and metatarsophalangeal stiffness. He presented pseudodislocation of the ankle joint which conditions talus-valgus feet. There was no resorption of the distal phalanges and he had normal nails. He had mild scoliosis, normal clavicles and no mandibular hypoplasia. He presented phlebomegaly in the limbs (L). Regarding comorbidities, he was diagnosed with diabetes mellitus, hypertriglyceridaemia, dilated cardiomyopathy with moderate pulmonary hypertension, for which he underwent a heart transplant, and he developed central nervous system lymphoma following the immunosuppressive therapy received. The images included here are of patients from our Lipodystrophy Unit (UETeM reference centre) and have not previously been published elsewhere.
Ijms 25 09324 g001
Figure 2. Body composition determined by Dual-Energy X-Ray Absorptiometry of patients with lipodystrophic laminopathies and a non-lipodystrophic subject. Colour mapped, total body composition scans via whole-body Dual-Energy X-Ray Absorptiometry of (A) a 42-year-old non-lypodystrophic woman, showing normal fat distribution; (B) a 48-year-old woman with classical FPLD2 (FPLD2 #1) and loss of fat in the upper and lower limbs and its accumulation in the face, neck and pubic area; (C) a 62-year-old woman with atypical FPLD2 (FPLD2 #2) showing loss of fat in the upper and lower limbs and less accumulation in the face and neck; (D) a 7-year-old boy with Hutchinson-Gilford progeria syndrome (HGPS #1), presenting with severe generalised lipodystrophy, including palms and soles as well as muscle atrophy; (E) a 16-year-old woman with Hutchinson-Gilford progeria syndrome (HGPS #2), presenting with severe generalised lipodystrophy, including palms and soles, and muscle atrophy; (F) a 17-year-old man with LMNA-atypical progeroid syndrome, showing generalised lipodystrophy, with preserved adipose tissue in the palms, loss of fat in the soles and muscle mass atrophy. Green represents an area of low level % fat (0–25%), yellow an area of medium level % fat (25–60%), and red an area of high level % fat (60–100%). These are original data from patients of our Lipodystrophy Unit (UETeM reference centre) and have not previously been published elsewhere. FPLD2: familial partial lipodystrophy type 2; HGPS: Hutchinson-Gilford progeria syndrome; APS: LMNA-atypical progeroid syndrome; BMI: body mass index; FM: fat mass; FFM: fat-free mass.
Figure 2. Body composition determined by Dual-Energy X-Ray Absorptiometry of patients with lipodystrophic laminopathies and a non-lipodystrophic subject. Colour mapped, total body composition scans via whole-body Dual-Energy X-Ray Absorptiometry of (A) a 42-year-old non-lypodystrophic woman, showing normal fat distribution; (B) a 48-year-old woman with classical FPLD2 (FPLD2 #1) and loss of fat in the upper and lower limbs and its accumulation in the face, neck and pubic area; (C) a 62-year-old woman with atypical FPLD2 (FPLD2 #2) showing loss of fat in the upper and lower limbs and less accumulation in the face and neck; (D) a 7-year-old boy with Hutchinson-Gilford progeria syndrome (HGPS #1), presenting with severe generalised lipodystrophy, including palms and soles as well as muscle atrophy; (E) a 16-year-old woman with Hutchinson-Gilford progeria syndrome (HGPS #2), presenting with severe generalised lipodystrophy, including palms and soles, and muscle atrophy; (F) a 17-year-old man with LMNA-atypical progeroid syndrome, showing generalised lipodystrophy, with preserved adipose tissue in the palms, loss of fat in the soles and muscle mass atrophy. Green represents an area of low level % fat (0–25%), yellow an area of medium level % fat (25–60%), and red an area of high level % fat (60–100%). These are original data from patients of our Lipodystrophy Unit (UETeM reference centre) and have not previously been published elsewhere. FPLD2: familial partial lipodystrophy type 2; HGPS: Hutchinson-Gilford progeria syndrome; APS: LMNA-atypical progeroid syndrome; BMI: body mass index; FM: fat mass; FFM: fat-free mass.
Ijms 25 09324 g002
Table 1. Differential features of Dunnigan disease in comparison with other familial partial lipodystrophy syndromes.
Table 1. Differential features of Dunnigan disease in comparison with other familial partial lipodystrophy syndromes.
Dunnigan Disease
[30,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87]
Other FPLD Subtypes
[88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110]
Molecular
characteristics
Nuclear lamina alteration
Variants in LMNA gene
- Classic FPLD2: R482W and R482Q variants in exon 8
- Atypical FPLD2: non-codon 482 variants
- Unknown: FPLD1.
- Adipogenesis dysregulation: FPLD3 (PPARG gene), FPLD9 (PLAAT3 gene).
- Lipid droplet impairment or lipolysis dysregulation: FPLD4 (PLIN1 gene), FPLD5 (CIDEC gene), FPLD6 (LIPE gene), FPLD8 (ADRA2A gene).
- Insulin signal transduction alteration: AKT2-related FPLD.
- Caveolar function alteration: FPLD7 (CAV1 gene).
- Dysregulation of phosphatidylcholine biosynthesis: PCYT1A-related FPLD.
- Mitochondrial dysfunction: MFN2-related FPLD.
Type of
inheritance
Autosomal dominant/Semi-dominant inheritance- Polygenic: FPLD1, FPLD7.
- Autosomal dominant: FPLD3, FPLD4, FPLD7, FPLD8 and AKT2-related FPLD.
- Autosomal recessive: FPLD5, FPLD6, FPLD9, PCYT1A- and MFN2-related FPLD.
Onset of fat lossPuberty in women, later in men- Birth: FPLD7
- Childhood: FPLD1, FPLD4, FPLD5, PCYT1A- and MFN2-related FPLD
- Adolescence: FPLD3, FPLD8 and MFN2-related FPLD
- Adulthood: FPLD3, FPLD4, FPLD6, AKT2- and MFN2-related FPLD
Abnormal
fat pattern
- Loss of fat in the limbs, trunk and gluteal region.
- Accumulation of fat in the face, neck, chin, axillae, interscapular area and abdominal viscera.
- Hypertrophy of mons pubis fat surrounded by subcutaneous lipoatrophy (“Dunnigan sign”).
- Subcutaneous lipomas (20%)
- FPLD3: less severe loss of fat.
- FPLD6 and MFN2-related FPLD: multiple lipomatous masses.
- FPLD7: loss of fat in the face and upper body.
- FPLD9: adipose tissue loss varies from partial to generalised.
Adipokine
Disturbance
- Leptin levels ranging from low to normal values.
- Lower adiponectin levels in comparison with healthy controls
MFN2-related FPLD: very low leptin concentrations.
Clinical features- Muscular hypertrophy
- Myalgias
- Phlebomegaly
- Hirsutism in women
- Acanthosis nigricans and acrochordons
- FPLD1: KöB index > 3.477
- FPLD3: less prominent musculature, phlebomegaly
- FPLD6: muscular dystrophy
- PCYT1A-related FPLD: short stature, muscular atrophy.
Organ abnormalities and comorbiditiesMetabolic abnormalities
- Non-ketotic diabetes
- Hypertriglyceridaemia, low HDL cholesterol
Heart abnormalities
- Cardiac hypertrophy
- Atrioventricular conduction defects
- Heart failure
- Early atherosclerosis
- Arrhythmias
Liver abnormalities
- NAFLD
- NASH
- Cirrhosis
Kidney abnormalities
- Proteinuria
- Chronic renal failure
Reproductive abnormalities
- Polycystic ovary syndrome
- Fertility problems
- Miscarriages and stillbirths
Others
- Hypertension
- Acute pancreatitis
- FPLD3: earlier and more severe metabolic complications
- FPLD5: diabetes with ketosis.
- FPLD6: auto-fluorescent drusen-like retinal deposits. Increased CK levels.
- FPLD7: congenital cataracts.
- FPLD9: neurological abnormalities (demyelinating neuropathy, intellectual disability).
- MFN2-related FPLD: peripheral axonal neuropathy.
FPLD: familial partial lipodystrophy; HDL: high-density lipoprotein; CK: creatine kinase; NAFLD: non-alcoholic fatty liver disease; NASH: non-alcoholic steatohepatitis.
Table 2. Differential diagnosis of the main lipodystrophy-associated laminopathies.
Table 2. Differential diagnosis of the main lipodystrophy-associated laminopathies.
DiseaseGeneInheritanceDysfunctionOnset of
Phenotype
Fat
Loss
Clinical FeaturesMain Comorbidities
FPLD2 or
Dunnigan
disease
[30,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87] #151660
LMNAADNuclear lamina
alteration
Puberty in women, later in menPartialSubutaneous lipomas.
Muscular hypertrophy, phlebomegaly.
“Dunnigan sign”
Diabetes, hypertriglyceridaemia, hepatic steatosis, fertility problems, PCOS, cardiovascular disease, arrhythmias.
May associate cardiomyopathy, muscular dystrophy.
HGPS
[118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137]
#176670
LMNAADNuclear lamina
alteration
18–24
months
GeneralisedGlobal alopecia, prominent veins on the scalp, micrognatia, dental crowding.Bone and joint alterations.
Atherosclerosis, cardiovascular disease.
Mild metabolic syndrome.
APS
[138,139,140,141,142,143,144,145,146,147]
-
LMNAADNuclear lamina
alteration
Childhood/Early adulthoodGeneralised or partialPhenotypic heterogeneity.
Premature greying of hair, alopecia, dental crowding.
Metabolic complications.
Valvulopathy, dilated cardiomyopathy.
Acroosteolysis and clavicular hypoplasia are absent or mild.
MADA
[8,14,159,160,161,162,163,164,165,166,167,168]
#248370
LMNAARNuclear lamina
alteration
2–4
years
PartialGrowth retardation, beaked nose, mottled skin pigmentation, dental crowdingBone alterations.
Metabolic syndrome.
Premature adrenal cortical dysfunction in some cases.
MADB
[131,157,169,170,171,172,173,174,175,176,177,178,179]
#608612
ZMPSTE24ARNuclear lamina
alteration
PerinatalGeneralisedShort stature, delayed closure of fontanels, calcified skin nodules.Musculoskeletal
abnormalities.
Metabolic syndrome.
NGPS
[9,12,180,181,182,183,184,185,186,187]
#614008
BANF1ARNuclear lamina
alteration
2 yearsGeneralisedProfound skeletal abnormalities.Bone alterations.
No metabolic complications.
FPLD2: familial partial lipodystrophy type 2; HGPS: Hutchinson Gilford progeria syndrome; APS: LMNA-atypical progeroid syndrome; MADA: Mandibuloacral dysplasia type A; MADB: Mandibuloacral dysplasia type B; NGPS: Nestor-Guillermo progeria syndrome; AD: autosomal dominant; AR: autosomal recessive; PCOS: polycystic ovary syndrome.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Díaz-López, E.J.; Sánchez-Iglesias, S.; Castro, A.I.; Cobelo-Gómez, S.; Prado-Moraña, T.; Araújo-Vilar, D.; Fernandez-Pombo, A. Lipodystrophic Laminopathies: From Dunnigan Disease to Progeroid Syndromes. Int. J. Mol. Sci. 2024, 25, 9324. https://doi.org/10.3390/ijms25179324

AMA Style

Díaz-López EJ, Sánchez-Iglesias S, Castro AI, Cobelo-Gómez S, Prado-Moraña T, Araújo-Vilar D, Fernandez-Pombo A. Lipodystrophic Laminopathies: From Dunnigan Disease to Progeroid Syndromes. International Journal of Molecular Sciences. 2024; 25(17):9324. https://doi.org/10.3390/ijms25179324

Chicago/Turabian Style

Díaz-López, Everardo Josué, Sofía Sánchez-Iglesias, Ana I. Castro, Silvia Cobelo-Gómez, Teresa Prado-Moraña, David Araújo-Vilar, and Antia Fernandez-Pombo. 2024. "Lipodystrophic Laminopathies: From Dunnigan Disease to Progeroid Syndromes" International Journal of Molecular Sciences 25, no. 17: 9324. https://doi.org/10.3390/ijms25179324

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