Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2018 Dec 12.
Published in final edited form as: Nat Rev Mol Cell Biol. 2017 Aug 9;18(10):595–609. doi: 10.1038/nrm.2017.68

Shared molecular and cellular mechanisms of premature ageing and ageing-associated diseases

Nard Kubben 1, Tom Misteli 1
PMCID: PMC6290461  NIHMSID: NIHMS999231  PMID: 28792007

Abstract

Ageing is the predominant risk factor for many common diseases. Human premature ageing diseases are powerful model systems to identify and characterize cellular mechanisms that underpin physiological ageing. Their study also leads to a better understanding of the causes, drivers and potential therapeutic strategies of common diseases associated with ageing, including neurological disorders, diabetes, cardiovascular diseases and cancer. Using the rare premature ageing disorder Hutchinson–Gilford progeria syndrome as a paradigm, we discuss here the shared mechanisms between premature ageing and ageing-associated diseases, including defects in genetic, epigenetic and metabolic pathways; mitochondrial and protein homeostasis; cell cycle; and stem cell-regenerative capacity.


Ageing is a process of gradual functional deterioration at the cellular and organismal level. About half of human deaths are attributed to chronic ageing-associated diseases (AADs), most prominently heart disease, diabetes, chronic obstructive pulmonary disease (COPD), stroke, Alzheimer disease, chronic kidney diseases (CKDs) and cancer1,2 (BOX 1). Despite ageing being the biggest risk factor for the development of these illnesses, our understanding of how the ageing process contributes to their onset and progression is rudimentary.

Box 1 | Chronic ageing-associated diseases.

Alzheimer disease

A chronic neurodegenerative disease characterized by dementia, disorientation, mood swings, loss of appetite, jumbled speech and an inability to coordinate movements; this disease is also associated with an increased risk for osteoporosis and muscle wasting. Familial forms of Alzheimer disease are mostly caused by a mutation in amyloid precursor protein (APP) and presenilins 1 and 2, which increase the production of a cleaved APP product called β‑amyloid (typically deposited in senile plaques). Furthermore, neurofibrillary tangles, consisting predominantly of hyperphosphorylated tau protein, are a hallmark of Alzheimer disease.

Atherosclerosis

A vascular disease characterized by arteries that are stiffened and calcified owing to a build‑up of cholesterol‑loaded plaques, which cause an obstruction of blood flow. Unstable plaques typically have decreased numbers of vascular smooth muscle cells and are more prone to rupture, which may cause heart attack or stroke.

Cancer

A group of diseases involving abnormal cell growth that manifests in either an invasive (malign) or a non‑invasive (benign) form as a result of accumulation of genetic mutations that either inhibit the activity of tumour suppressor genes, or activate or overexpress oncogenes.

Chronic kidney disease

A chronic condition typified by a gradual loss of kidney function over time, which can result in high blood pressure, anaemia, loss of bone mass and neuronal damage.

Chronic obstructive pulmonary disease

A group of lung diseases that reduce airflow, cause difficulty breathing and predominantly result from either increased mucus production and inflammation (bronchitis) or the destruction and enlargement of the air spaces (emphysema). Idiopathic pulmonary fibrosis is characterized by a thickening and scarring of the lung, which reduces the exchange of oxygen with the bloodstream. Patients with chronic obstructive pulmonary disease (COPD) are at increased risk for the development of Parkinson disease.

Heart failure

A permanent state of insufficient cardiac output owing to the failure of the heart to properly contract or relax as a result of stretching/thinning or hypertrophy/stiffening, respectively, of the ventricular walls.

Osteoporosis

A loss of bone mass owing to an imbalance between the bone formation and bone resorption processes.

Parkinson disease

A long‑term neurodegenerative disorder affecting motor system function that results in shaking, rigidity, difficulty walking and, in many cases, dementia and depression. Hereditary forms of Parkinson disease include mutations in α‑synuclein, parkin, leucine‑rich repeat serine/threonine‑protein kinase 2 (LRRK2), PTEN‑induced putative kinase protein 1 (PINK1), DJ1 and ATP13A2. A typical hallmark of Parkinson disease includes the accumulation of α‑synuclein in the form of Lewy bodies, which contribute to cell death in the dopaminergic‑innervated substantia nigra.

Sarcopenia

A degenerative loss of muscle mass and quality with ageing. Patients with COPD, heart failure, cancer or chronic kidney disease have an increased occurrence of sarcopenia.

Type 2 diabetes

A metabolic disorder in which insulin is not used properly, which is initially compensated for by increased production of insulin by pancreatic β‑islet cells. Ultimately, the β‑islet cells fail. Patients with type 2 diabetes are at risk for the development of Alzheimer disease as well as for chronic kidney failure.

The role of ageing in human disease is commonly studied in animal disease models by delaying the onset and progression of ageing-associated defects in the tissue that is primarily associated with the various ageing-related diseases (BOX 1). Although animal models are a useful surrogate to study the fundamentals of ageing, which may be conserved across species, these systems are not ideal to elucidate the effects of ageing on human disease owing to the much lower frequency at which chronic AADs occur in laboratory animals compared with humans35.

A key contributor to AADs is the ageing-related decline of cell and tissue function6,7. Cellular ageing is characterized by increased genomic instability, altered metabolism and the loss of regenerative potential. Cellular ageing and deterioration as a driver in AADs explain the observation that in many AADs, not only the tissue primarily associated with the disease is affected, but other tissues simultaneously undergo functional decline6,7. The often overlooked functional decline across multiple organs in AADs is important for disease pathology and diagnosis, as non-primary tissue defects can be used as an independent disease predictor: for example, handgrip strength and hip fractures are indicators of CKD and preclinical Parkinson disease, respectively8,9. These observations indicate that cellular ageing is a universal principle underlying AADs.

The study of human ageing has been facilitated by the discovery of mutations that cause premature ageing syndromes (TABLE 1). Most prominent among them are Hutchinson–Gilford progeria syndrome (HGPS)1012 and atypical Werner syndrome, which are caused by genetic defects in nuclear envelope proteins, as well as classical Werner syndrome, Cockayne syndrome, Bloom syndrome, xeroderma pigmentosum, ataxia telangiectasia, trichothiodystrophy, dyskeratosis congenita (DKC) and mosaic variegated aneuploidy syndrome, which are caused by defects in DNA repair and maintenance proteins (TABLE 1). The cellular defects that are observed in these and other human premature ageing models, including genomic and proteomic instability, altered metabolism and loss of regenerative potential, overlap with defects that occur during physiological ageing in humans. Moreover, there are striking commonalities between organismal defects in several premature ageing diseases and AADs. However, as premature ageing syndromes only represent some aspects of the normal ageing process, it is not surprising that progeroid syndromes are typically only associated with a subset of AADs (TABLE 1). For example, HGPS involves extensive cardiovascular and osteoporotic pathology, whereas premature ageing disorders caused by mutations in DNA repair proteins are often characterized by cancer susceptibility and neurodegeneration. These parallels suggest common aetiologies between premature ageing syndromes and AADs.

Table 1 |.

Progeroid syndromes

Progeroid syndrome Gene Main function of gene Symptoms Non-represented chronic ageing diseases (BOX 1) Parallel with leading chronic ageing diseases (BOX 1)
Hutchinson-Gilford progeria syndrome LMNA Nuclear envelope architectural protein, heterochromatin organization Alopecia, atherosclerosis, growth retardation, loss of subcutaneous fat, skeletal muscle wasting, nail dystrophy, stiff joints, tight skin, subcutaneous calcifications, osteoporosis, loss of eyesight, kidney failure Cancer, CKD, COPD, diabetes, neurodegenerative diseases Atherosclerosis, heart failure, skeletal muscle wasting, osteoporosis
Nestor-Guillermo progeria syndrome BANF1 Nuclear envelope architectural protein, heterochromatin organization Alopecia, growth retardation, loss of subcutaneous fat, skeletal muscle wasting, stiff joints, osteoporosis Atherosclerosis, cancer, CKD, COPD, diabetes, heart failure, neurodegenerative diseases Osteoporosis, skeletal muscle wasting
Werner syndrome WRN (classical), LMNA (atypical) WRN: DNA repair (NER, BER, NHEJ, HR), telomere maintenance Growth retardation, tight skin, skin ulcers, osteoporosis, cataract, cardiac valve and soft tissue calcification, loss of subcutaneous fat, decreased fertility, increased risk for cancer Cancer, CKD, COPD, diabetes, neurodegenerative diseases, skeletal muscle wasting, diabetes Atherosclerosis, increased risk for cancer, heart failure, osteoporosis
Cockayne syndrome ERCC6, ERCC8 DNA repair (NER) Growth retardation, impaired development of the nervous system and progressive neurodegeneration, photosensitivity, cataracts Atherosclerosis, cancer, CKD, COPD, diabetes, heart failure, osteoporosis, skeletal muscle wasting Progressive neurodegeneration
Bloom syndrome BLM Double-strand break DNA repair Growth retardation, photosensitivity, micrognathism, skin rash, dilated blood vessels, moderate immune deficiency, cancer, increased risk diabetes and COPD Atherosclerosis, CKD, heart failure, osteoporosis, neurodegenerative diseases, skeletal muscle wasting Cancer, increased risk for diabetes and COPD
Xeroderma pigmentosum XPA, XPB, XPC, XPD, XPG, ERCC4, ERCC6, DDB2, POLH, RAD2 DNA repair (NER) Photosensitivity, cancer, dilated blood vessels Atherosclerosis, CKD, COPD, diabetes, heart failure, osteoporosis, neurodegenerative diseases, skeletal muscle wasting Cancer
Ataxia telangiectasia ATM DNA damage signaling activator Growth retardation, weakened immune system, cancer, degeneration of cerebellum, dilated blood vessels, diabetes Atherosclerosis, CKD, COPD, diabetes, heart failure, osteoporosis, skeletal muscle wasting Cancer, diabetes, neuronal degeneration
Trichothiodystrophy XPB, XPD, TFB5 DNA repair (NER) Growth retardation, brittle hair, nail dystrophy, intellectual impairment, neuronal degeneration, reduced fertility Atherosclerosis, cancer, CKD, COPD, diabetes, heart failure, osteoporosis, skeletal muscle wasting, diabetes Neuronal degeneration
Dyskeratosis congenita TERC, TERT, CTC1, WRAP53 Components of telomerase and telomere maintenance complex Nail dystrophy, leukoplakia of oral mucosa, bone marrow failure, hyperpigmentation of skin, premature greying of hairs, testicular atrophy, cancer, osteoporosis, pulmonary fibrosis Atherosclerosis, CKD, diabetes, heart failure, neurodegenerative diseases, skeletal muscle wasting Cancer, osteoporosis, pulmonary fibrosis
Mosaic variegated aneuploidy syndrome BUB1B, CEP57 Mitotic non-disjunction Short stature, central nervous system and brain abnormalities, intellectual disability, aneuploidy, increased cancer risk, cataracts Atherosclerosis, CKD, COPD, diabetes, heart failure, osteoporosis, neurodegenerative diseases, skeletal muscle wasting Cancer

BER, base excision repair; CKD, chronic kidney disease; COPD, chronic obstructive pulmonary disease; HR, homologous recombination; NER, nucleotide excision repair; NHEJ, non-homologous end joining.

In this Review, we discuss defects in cell and molecular mechanisms that are common between human premature ageing diseases and physiological ageing, and we highlight the role of these pathways in several human AADs. We use HGPS as a paradigm for the interplay between premature ageing, physiological ageing and AADs because it is one of the best characterized premature ageing diseases and exhibits a wide range of cellular, tissue and organismal defects that are shared with AADs.

HGPS as a Rosetta Stone for ageing mechanisms

HGPS is an ultra-rare premature ageing disease that affects 1 in 4–8 million newborns13. The clinical features of HGPS become overt shortly after birth and include full-body alopecia, loss of subcutaneous fat and skeletal muscle, nail dystrophy, joint stiffening, wrinkling of the skin, subcutaneous calcifications, weakening of the bone structure and loss of eyesight13 (TABLE 1). HGPS is invariably fatal, with atherosclerosis-resembling progressive cardiovascular disease, myocardial infarction and stroke being the most common causes of death during the patient’s mid-teens14. HGPS is predominantly caused by a heterozygous silent mutation (G608G) in the LMNA gene, which leads to the production of progerin, an aberrant form of the nuclear structural protein lamin A10,11,15. Wild-type lamin A is post-translationally processed: cleavage of the C terminus by CAAX prenyl protease 1 homologue (encoded by ZMPSTE24, also known as ZMPSTE24 in mice) releases mature lamin A, which then undergoes higher-order assembly into intermediate filaments that are incorporated in the nuclear lamina and the nuclear interior16,17 (FIG. 1). In HGPS, the 1824C>T mutation (LMNAG608G) activates a cryptic splice donor site, leading to the removal of a 50-amino-acid region containing the ZMPSTE24 recognition site. As a result, progerin accumulates at the nuclear periphery and disrupts the mechanochemical properties of the nuclear lamina, as indicated by the aberrant nuclear shape seen in cells from patients with HGPS17 (FIG. 1). Similar cellular defects are caused by LMNA mutations that cause atypical HGPS, including LMNAG608S and LMNAE145K (REFS 16,18). Defects in mechanical properties and mechanotransduction at the nuclear lamina are thought to be relevant to the pathology of patients with HGPS, as many of the affected tissues, such as the vasculature, bone and joints, are exposed to mechanical forces17,19.

Figure 1 |. Cellular ageing defects.

Figure 1 |

An overview of the cellular ageing defects that underlie premature ageing disorders, regular ageing and ageing-associated disease. These ageing defects are interconnected and include reduced DNA repair pathway efficiency, loss of genomic integrity, a global loss of heterochromatin, alterations in metabolic signalling, increased formation of reactive oxygen species (ROS) by mitochondria, reduced activity of proteostasis-promoting proteolytic pathways and activation of senescence pathways. Post-translational processing of wild-type prelamin A by the zinc metalloproteinase ZMPSTE24 releases mature wild-type lamin A that is incorporated in the nuclear lamina (indicated in the left half of the cell nucleus). In the premature ageing disease Hutchinson–Gilford progeria syndrome (HGPS), the LMNAG608G mutation activates a cryptic splice site that results in the formation of a prelamin A isoform lacking 50 amino acids, including the ZMPSTE24 cleavage site, and is named progerin (indicated in the right half of the cell nucleus). Incorporation of this mutant isoform into the nuclear lamina distorts the nuclear shape and further contributes to other ageing defects shown in the figure (thin arrows indicate dysregulated pathways and mechanisms). AMPK, AMP-activated protein kinase; HR, homologous recombination; IL-1, interleukin-1; NER, nucleotide excision repair; NHEJ, non-homologous end joining; NF-κB, nuclear factor-κB; PGC1α, proliferator-activated receptor-γ co-activator 1α; SIRT1, sirtuin 1.

HGPS differs from other human premature ageing syndromes by its early onset, severity of ageing symptoms and wide range of affected tissues15,20 (TABLE 1). In contrast to HGPS, the classical adult progeria Werner syndrome is characterized by the onset of osteoporosis and cancer by the third decade of life. The early emergence of ageing defects in patients with HGPS can be partially attributed to the dominant-negative mode of action of progerin, whereas other progeroid syndromes are often associated with recessive genetic defects that result in a loss of function of DNA repair proteins, which might be partially compensated by alternative DNA repair pathways20. Consequently, patients with HGPS have a wider range of tissue defects, from osteoporosis to skeletal muscle- and cardiovascular-related ageing pathologies, whereas DNA repair-associated progeroid syndromes typically involve higher cancer predisposition and, in some cases, progressive neurodegeneration20 (TABLE 1). Interestingly, for unknown reasons, patients with HGPS are not at increased risk of developing diabetes and chronic kidney failure. Moreover, they do not exhibit neurodegeneration, probably as a result of microRNA-9 (miR-9) repressing the neuronal expression of lamin A and progerin. Furthermore, despite the high levels of DNA damage, patients with HGPS do not develop childhood tumours, which could result from the altered composition of the extracellular matrix reducing the invasion of tumour-initiating cells21,22 and progerin exhibiting a tumour-protective function23.

Although several AAD symptoms are absent in HGPS (TABLE 1), many of the hallmarks of cellular ageing observed in tissues affected in AADs are also seen in patients with HGPS24 (FIG. 1). This prominent display of cellular ageing defects in HGPS has been attributed to the detrimental effects of progerin on the nuclear lamina, which is a key organizer of the mammalian cell nucleus25. The lamina has been mechanistically linked to most of the universal hallmarks of ageing and AADs (FIG. 1), including loss of genomic and epigenetic integrity, shortened telomeres, reduced protein homeostasis (proteostasis), metabolic reprogramming, mitochondrial dysfunction, cellular senescence and reduced stem cell maintenance and regenerative capacity26. The expression of progerin is thus implicated in diverse pathways, whereas progeroid syndromes like Werner syndrome, Bloom syndrome and Cockayne syndrome have a more limited range of defects that originate mainly from defective DNA repair2734. The importance of the nuclear lamina in ageing is further illustrated by the premature ageing syndrome Néstor–Guillermo progeria syndrome (NGPS), which is caused by a homozygous mutation in the nuclear lamina-localized protein barrier-to-autointegration factor (BANF1)35. Similar to HGPS, NGPS is associated with joint stiffness, loss of subcutaneous fat and skeletal muscle, and growth retardation35 (FIG. 1). The implication of nuclear lamina dysfunction in various hallmarks of cellular ageing makes HGPS a suitable model to discuss the parallels between premature ageing and AADs at the cellular and molecular level.

HGPS is especially relevant to normal ageing because physiologically aged individuals also generate small amounts of progerin owing to the spontaneous sporadic usage of the cryptic splice site that is activated in patients with HGPS28,36. Given the dominant-negative action of progerin, these low levels of progerin are thought to contribute to normal ageing36. This view is supported by extensive similarity of cellular and organismal defects in patients with HGPS and normally aged individuals20,26. In addition, genetic abrogation of progerin and lamin A expression extends the lifespan of mice37, and the LMNA 1968G>A mutation, which results in lower activation of the cryptic splicing event of progerin, results in a milder ageing-related pathology38.

The single, well-defined genetic defect in patients with HGPS, the direct links between progerin and numerous cellular ageing hallmarks, the expression of the disease-causing protein during normal physiological ageing and the manifestation of ageing-related pathologies across multiple tissues make HGPS an attractive model to study the role of ageing mechanisms in AADs.

Genomic instability in ageing and AADs

The human genome is under continuous attack by DNA-damaging insults that threaten cellular homeostasis. These assaults are counteracted by mechanisms that repair DNA damage and protect chromosomal ends from shortening, which decline during human ageing3942 and in AADs, leading to cellular deterioration.

Defects in DNA repair pathways

Several DNA damage repair pathways maintain the integrity of the genome. Nucleotide excision repair (NER) pathways repair single-strand DNA damage4042, whereas double-strand breaks (DSBs), which promote large chromosomal rearrangements that threaten cell survival, are preferentially and faithfully repaired by homologous recombination but can also be repaired by error-prone non-homologous end joining (NHEJ). The efficiency of these pathways decreases during human ageing3942 (FIG. 2), resulting in increased chromosomal aberrations and permanent activation of DNA damage signalling with age43.

Figure 2 |. Defects in DNA damage repair associated with ageing.

Figure 2 |

An overview of DNA repair pathways that are impaired in premature ageing syndromes, ageing and ageing-associated diseases. Thin arrows indicate the DNA repair proteins that are impaired specifically in Hutchinson–Gilford progeria syndrome (HGPS). Double-strand breaks are either repaired by homologous recombination through RAD51-mediated repair using the sister chromosome as a template, or X-ray repair cross-complementing protein 6 (XRCC6)- and DNA protein kinase catalytic subunit (PKcs)-mediated non-homologous end joining of the broken ends. Poly(ADP-ribose) polymerase 1 (PARP1) promotes ataxia telangiectasia mutated (ATM)-dependent DNA damage signalling activation. In nucleotide excision repair, bulky DNA adducts are excised, and the resulting single-strand DNA breaks are repaired by the indicated proteins. The final step of repair for each repair pathway includes DNA ligase-mediated joining of the DNA strands at the breakage point. PCNA, proliferating cell nuclear antigen.

DNA repair is impaired in HGPS.

In HGPS, chronic activation of DNA damage signalling results from an increased number of replication-induced DSBs and their defective repair44. DNA replication defects, including replication fork collapse, are a prominent source of DSBs and occur more frequently in HGPS because of reduced levels of the critical replication DNA clamp protein PCNA, disrupted interactions between lamin A and proliferating cell nuclear antigen (PCNA), and increased proteolytic degradation of the replication factor C1 (REFS 4547) (FIG. 2). Homologous recombination and NHEJ-mediated repair of these DSBs is impaired by the delayed recruitment of the repair proteins TP53-binding protein 1 (TP53BP1) and RAD51, inhibition of the poly(ADP-ribose) polymerase 1 (PARP1) DNA repair protein, and reduced levels of the NHEJ-associated proteins DNA-PKcs and X-ray repair cross-complementing protein 6 (XRCC6)30,48,49. Moreover, aberrant sequestration of the NER repair protein XPA at sites of DSBs further interferes with DSB repair and decreases NER45 (FIG. 2).

Impaired DNA repair in other progeroid syndromes.

A causal role of NER and DSB repair in ageing and AADs is further suggested by the progeroid phenotypes in Cockayne syndrome, xeroderma pigmentosum and trichothiodystrophy, all of which are caused by defects in NER proteins, as well as in ataxia telangiectasia, Bloom syndrome and Werner syndrome, all of which are defective in DSB repair20 (FIG. 2; TABLE 1). The increased prevalence of neuronal pathologies among patients with some of these progeroid syndromes indicates that postmitotic neurons crucially depend on DNA repair efficiency, probably because they accumulate DNA damage during the organismal lifespan as a result of high transcriptional activity and exposure to oxidative metabolic stress50. Paradoxically, the efficiency of the DNA repair system in the brain is low, and it is even lower in neurodegenerative diseases50. This observation is consistent with the increased incidence of aneuploid neurons in Alzheimer disease50,51, which probably results from defective NHEJ-mediated repair and is consistent with the finding that mutations of the parkin protein in familial Parkinson disease52 (BOX 1) directly inhibit NER. Moreover, both these diseases cause fibroblasts and lymphoblasts to become hypersensitive to X-ray-induced DNA damage50,53. These increased amounts of DNA damage may functionally impair neurons and promote an apoptosis-driven decline in neuronal mass50.

DNA repair impairment in AADs.

Increased DNA damage is also seen in the specific cell types that are affected in AADs. Vascular smooth muscle cells (VSMCs), which become defective in atherosclerosis, have higher amounts of DNA damage after exposure to reactive oxygen species (ROS), owing to increased mechanical stress, lipoprotein accumulation and inflammatory responses. The subsequent reductions in the proliferation and apoptosis of VSMCs may promote atherosclerotic plaque rupture54. VSMCs become increasingly senescent in patients with HGPS in conjunction with defective homologous recombination repair48. Similarly, lymphocytes from patients with type 2 diabetes (T2D) have reduced capacity to repair ROS-induced DNA damage, and NHEJ deficiency impairs pancreatic β-cells55,56. Furthermore, differentiation of bone-forming osteoblasts is inhibited by DNA damage and contributes to weakened bone structure in a progeria mouse model57. In line with this evidence, defective NHEJ can result in osteopenia, a less severe ageing-related form of osteoporosis58.

Decreased DNA repair efficiency leads to the accumulation of genomic mutations during ageing, which promotes tumorigenesis through activation of oncogenes or inactivation of tumour suppressor genes. This mechanism is exemplified by the increased incidence of tumours observed in most of the DNA repair pathway-associated progeroid syndromes20 (TABLE 1).

The above examples suggest that DNA damage is a key factor in AADs, but a key question that remains open is whether the effect of DNA repair on genome integrity is stochastic or affects specific genome regions59.

Reduced telomere length

Telomeres are repetitive sequences at the distal ends of chromosomes that are capped by the shelterin complex, which includes the telomeric repeat-binding factor 1 (TERF1) and TERF2 proteins, and protect chromosomal ends from being recognized as DSBs60. Telomeres shorten during cell division (a process known as telomere attrition) and can trigger DNA damage signalling and cellular senescence when they reach a critically short length. Telomerase, a complex consisting of telomerase reverse transcriptase (TERT) and telomerase RNA component (TERC), which elongates telomeres at each cell cycle, is highly expressed in embryonic stem cells; however, it is undetectable in most other human cells61.

Patients with HGPS show increased activation of DNA damage signalling at telomeres; their telomere length is reduced62,63 and telomeric chromosomal aberrations increased owing to impaired homologous recombination64. Similarly, chromosomal end-to-end fusions occur in the progeroid Terc-knockout mouse model20. Progerin may further increase telomere damage by disrupting both endogenous lamin A and TERF2-dependent protective stabilization of telomere ends, which is also affected in patients with atypical Werner syndrome6568.

Telomere damage can trigger the synthesis of progerin in wild-type cells, and the loss of function of the telomerase complex causes the premature ageing syndrome DKC, which suggests a causal role for telomere shortening in AADs20,69 (TABLE 1). Genetic ablation of Terc RNA in mice shortens telomeres to a critical length within a few generations, leading to cardiac myocyte hypertrophy, impaired left ventricular function and increased systolic blood pressure, all of which are normally associated with ageing61. Additionally, Terf2-knockout mice show increased apoptosis of cardiac myocytes60. A role for telomere shortening in cardiovascular disease is further supported by Terc-knockout mice presenting an atherosclerotic phenotype, and by vascular endothelial cells and VSMCs from human patients who have increased haemodynamic stress and atherosclerotic regions having shorter telomeres60. Increasing telomerase activity can be beneficial because it prolongs the lifespan of vascular endothelial cells, but it is also detrimental because it promotes neointimal VSMC and leukocyte proliferation, which aggravates the atherosclerotic phenotype60.

Telomere attrition also promotes osteoporosis in patients with DKC. Chondrocytes and peripheral blood leukocytes in elderly individuals with osteoporosis have shorter telomeres, and restoring telomerase activity in osteoblasts increases bone deposition70. Mutations that result in TERT loss of function and telomere attrition also cause ageing-associated lung emphysema and fibrosis71. Studies in mouse models suggest that telomere attrition accelerates the development of ageing-related lung emphysema by reducing the ability of the lungs to withstand toxin-induced stress72,73. In CKD, telomere attrition was found to reduce cell viability and increase renal injury upon ischaemia–reperfusion71. The same relationship may hold for pancreatic β-cells, the survival and functionality of which are compromised under high blood sugar levels in Tert-knockout mice71.

Although TERT expression in fibroblasts from patients with HGPS can rescue certain cellular ageing-associated defects, the observation that progerin and TERT-expressing human fibroblasts retain some ageing defects suggests that other cellular ageing mechanisms may either regulate the integrity of telomeres and other genomic regions or drive ageing32,74.

Epigenetic defects

Chromatin structure is modulated by DNA methylation and histone post-translational modifications, which affect genomic integrity, gene expression and ultimately cellular health and disease26,75.

Changes in histone methylation in ageing

Cells from aged individuals show a global loss of his-tones and a progressive loss of histone H3 trimethylation at lysines 9 and 27 (H3K9me3 and H3K27me3), which are repressive marks that promote chromatin compaction26 (FIG. 3). Loss of heterochromatin is increased by ageing-associated downregulation of heterochromatin protein 1 homologue-α (HP1α; also known as CBX5), the nucleosome remodelling and deacetylase (NuRD) chromatin remodelling complex and Polycomb group (PCG) proteins, all of which are epigenetic silencers20,32,76. Global loss of heterochromatin, HP1α and NuRD complex protein members has also been reported in cells from patients with HGPS28,77. In these cells, loss of the NuRD complex and decreased expression of the H3K27-specific methyltransferase EZH2 results in the depletion of the repressive H3K27me3 heterochromatin mark32,77 (FIG. 2). Interestingly, knockdown of the H3K27 demethylase UTX1 in Caenorhabditis elegans extends the lifespan by 30%78. Decreased activity of the Polycomb repressive complex 2 (PRC2) member EZH2 has been implicated in T2D, as conditional deletion of EZH2 in pancreatic β-cells in juvenile mice reduces β-cell proliferation and mass through induced cellular senescence, resulting in a diabetic phenotype79.

Figure 3 |. Epigenetic defects associated with ageing.

Figure 3 |

An overview of the epigenetic alterations (indicated by thin arrows) on the histone H3 tail that contribute to global loss of heterochromatin in Hutchinson–Gilford progeria syndrome (HGPS) and are associated with ageing and ageing-associated diseases. Decreased levels of EZH2 reduce the trimethylation (indicated by the 3 green hexagons) of histone H3 at lysine 27 (H3K27) by the Polycomb repressive complex 2 (PRC2) Polycomb group (PCG) protein complex77, which is a repressive chromatin mark that enables binding of the PRC1 PcG complex and the subsequent trimethylation (H3K9me3), which can bind heterochromatin protein 1 homologue-α (HP1α); these methylation levels are decreased in HGPS28. H3K9me3 has been reported to be downregulated in cells from patients with HGPS but was conversely found to be upregulated in a Zmpste24-knockout progeria-like mouse model28,83. Various proteins within the nucleosome remodelling and deacetylase (NuRD) complex have decreased expression levels in HGPS32, which reduces the histone deacetylation activity of this complex. H3K9 and H3K27 acetylation (indicated by red triangles) are expression-permissive chromatin marks that are thought to be mutually exclusive to the repressive trimethylation marks on the same lysines. HDAC1, histone deacetylase 1; RBB4, RB-binding protein 4.

The major function of the PRC2 complex is to establish H3K27me3 marks, to which the PRC1 complex binds, and to induce transcriptional silencing by promoting histone lysine N-methyltransferase SUV39H1-mediated H3K9me3 (REF. 80) (FIG. 3). Disruptions of the nuclear lamina affect PCG protein localization, resulting in genome-wide loss of H3K9me3 in HGPS; this loss seems to contribute to defective telomere maintenance and aberrant activation of pericentromeric satellite repeats, which are normally silent32,77,81. Progressive loss of HP1α, which binds to H3K9me3, may further contribute to hetero chromatin depletion, which cannot be compensated for by the globally increased levels of the H4K20me3 heterochromatin mark77. Progressive loss of H3K9me3 also occurs in human cells lacking the Werner protein, probably as a result of reduced SUV39H1 activity and HP1α binding82. Conversely, increased levels of H3K9me3 are observed in a progeria-like mouse model in which the gene for the prelamin A-processing enzyme ZMPSTE24 is deleted. In this mouse model, SUV39H1 depletion restores genomic integrity and increases lifespan83. Loss of H3K9 methylation has also been linked to several AADs, and H3K9me3 depletion activates a vascular inflammatory gene signature in CKD84. In familial Alzheimer disease, the accumulation of mutant tau protein reduces H3K9me2 and HP1α levels, which results in upregulation of the neurotoxic piwi-like protein 1 (PIWIL1)85. Moreover, reduced expression of HP1α occurs in multiple cancers and correlates with poor prognosis86.

Ageing affects histone acetylation

The global loss of histone methylation in HGPS is accompanied by hypoacetylation of histones H2B and H4, possibly owing to diminished association of the histone acetyltransferase KAT8 with the nuclear lamina87. Interestingly, either KAT8 overexpression or treatment with histone deacetylase (HDAC) inhibitors extends the lifespan of progeria-like Zmpste24-knockout mice87. The H4K16 acetylation (H4K16ac) mark, which is associated with both transcriptional activation and repression, promotes NHEJ and homologous recombination, and increased levels of H4K16ac are associated with an extended cellular lifespan88,89. Treatment with HDAC inhibitors can prevent ageing-associated cognitive decline and reduce the severity of ischaemic stroke, Parkinson disease and osteoporosis in mouse models of disease by altering the expression of disease-promoting and disease-repressing genes90,91. HDACs, including sirtuins, also regulate the acetylation of non-histone proteins, including master regulators of cell growth and metabolism, suggesting a direct role for epigenetic regulation in the metabolic control of ageing26.

Cellular metabolism

Metabolic signalling is crucial for the maintenance of cellular homeostasis, as such signalling allocates energy to essential protective mechanisms that safeguard the integrity of the genome, epigenome, proteome and cellular organelles. Dysregulated metabolic signalling, which triggers a cascade of deleterious cellular events, is widely implicated in AADs.

Nutrient-sensing metabolic pathways

Maintaining a proper balance between catabolic and anabolic pathways is essential to maintaining adequate cellular energy levels. Disruption of this equilibrium perturbs cellular homeostasis, and drives cellular ageing and AADs26. Nutrient-sensing signalling pathways, including the insulin and insulin-like growth factor 1 (IGF1) signalling (IIS) pathway, and the sirtuin 1 (SIRT1) and AMP-activated protein kinase (AMPK) pathways, regulate the metabolic status of cells and have prominent roles in physiological ageing26.

Dysregulated IGF1 signalling causes ageing.

The IIS pathway is activated when there is an excess of nutrients, including insulin, IGF1 and free amino acids. This activity results in downstream activation of mTOR (FIG. 4) and, through various downstream effectors, increases protein synthesis and the promotion of other anabolic processes92. Reduced IIS signalling, either caused by genetic polymorphisms or following caloric restriction, promotes longevity and healthspan26. However, these effects of reduced IIS signalling are dependent on the age of the organism, the duration and the degree of repression, and the levels of cellular stressors such as inflammation. While short and moderate IIS repression may be beneficial by minimizing cellular growth to direct energy towards the repair of cellular damage, chronically increased attenuation is deleterious and promotes ageing92. In the progeria-like Zmpste24-knockout mouse model, IGF1 levels are drastically reduced from a neonatal age throughout life, and IGF1 treatment extends lifespan and delays the manifestation of progeroid features93. Parkinson disease is characterized by prolonged attenuation of mTOR, which results in increased neuronal death; this activity in animal models can be partially prevented by mTOR activation94.

Figure 4 |. Mitochondrial ROS-driven ageing defects.

Figure 4 |

Mitochondrial oxidative phosphorylation in ageing and ageing-associated diseases (AADs) is driven by a decrease in insulin-like growth factor 1 (IGF1) signalling, or uncoupling of IGF1 and mTOR signalling resulting from chronic activation of the IGF-1 pathway (indicated by interruption of arrow), and increased activation of mTOR, resulting from impaired AMP-activated protein kinase (AMPK) and sirtuin 1 (SIRT1) signalling26. Increased formation of reactive oxygen species (ROS) due to increased mitochondrial activity, decreased levels of antioxidants and decreased PCG1α (peroxisome proliferator-activated receptor-γ co-activator 1α) activity, resulting in impaired mitochondrial biogenesis and turnover, causes damage to DNA, proteins and other macromolecules27. Decreased efficiency of DNA repair pathways and proteostasis pathways, including autophagy and ubiquitin–proteasome system (UPS)-mediated degradation of damaged proteins, during (premature) ageing and in AADs further contributes to the deleterious effects of ROS on mitochondrial integrity and cellular homeostasis26.

Sustained IIS activation can also be detrimental to cellular health (FIG. 4). T2D-induced chronic mTOR activation deactivates insulin receptor substrate 1 (IRS1) through a feedback loop that uncouples insulin and IGF1 receptors from downstream signalling and causes insulin resistance94. This mechanism causes diabetes-associated symptoms in patients with Werner syndrome that can be alleviated with mTOR inhibitors20,95. Neurons are highly sensitive to alterations in IIS signalling because of their elevated metabolic rate and dependency on glucose94. In patients affected by Alzheimer disease, symptoms of diabetes — which result from uncoupling and inefficient IGF1 receptor signalling, and are aggravated by inhibiting insulin receptors with β-amyloids — precede cognitive decline by decades. Treatment with the mTOR inhibitor rapamycin partially restores insulin–mTOR signalling and alleviates neuronal defects in animal models of Alzheimer disease94,96.

Decreased sirtuin and AMPK activity contributes to AADs.

The protein deacetylase SIRT1 stimulates pathways that increase cellular energy levels and promote cell survival under conditions of stress and increased levels of NAD+ (REF. 26). During ageing, increased DNA damage signalling depletes NAD+, reducing SIRT1 activity97 (FIG. 4). The importance of decreased SIRT1 activity in ageing is supported by the finding that its restoration in progeria-like Zmpste24-knockout mice, resulting from treatment with the SIRT1 activator resveratrol, improves osteoporotic pathologies and extends lifespan98. SIRT1 activation also alleviates symptoms of Alzheimer disease, as it inhibits mTOR, which is hyper-activated in Alzheimer disease, and induces mTOR-independent anti-amyloidic cleavage and degradation of tau protein92 (TABLE 1).

SIRT1 activation also increases healthspan through activation of AMPK. In response to elevated AMP:ATP ratios, AMPK inhibits mTOR, promotes lipid catabolism and gluconeogenesis and, through positive feedback, activates SIRT1 (REF. 99) (FIG. 4). Like SIRT1, AMPK activity decreases with age. Consistent with this decline, treatment with the AMPK activator metformin extends the lifespan of C. elegans and mice, reduces kidney fibrosis in individuals with CKD and exerts anti-diabetic effects26,100. T2D metabolic defects are also ameliorated by increased activation of SIRT6, possibly through suppression of IGF1 signalling92. SIRT6 levels are decreased in patients with HGPS, and SIRT6 over-expression partially rescues the senescent phenotype of fibroblasts from patients with HGPS and extends the lifespan of wild-type male mice101,102.

Although it is evident that ageing is linked to IIS, AMPK and sirtuin-mediated metabolic signalling, how the downstream activators of these pathways affect cellular ageing in disease is only partially understood. An important factor that impacts cellular ageing in this context is the metabolic control of mitochondrial integrity.

Mitochondrial dysfunction and oxidative damage

Mitochondria produce ATP in an oxygen-dependent manner through oxidative phosphorylation (OXPHOS) by creating a proton gradient across the inner mitochondrial membrane (through electron transport chain protein complexes), which is then used by ATP synthase. In addition, mitochondria regulate gluconeogenesis, fatty acid oxidation, intracellular calcium levels and apoptosis103.

Decreased OXPHOS activity in HGPS, physiological ageing and AADs.

Mitochondrial integrity becomes compromised during ageing, as indicated by decreased transmembrane potential, increased formation of ROS at the expense of ATP synthesis, calcium dysregulation, apoptosis induction and increased mutations in mitochondrial DNA (mtDNA, which encodes proteins in the OXPHOS system)103. The connection between mitochondrial integrity and ageing is demonstrated by knock-in mice that express a proofreading-deficient version of the mtDNA polymerase POLGα. In these mice, OXPHOS efficiency is reduced owing to the accumulation of mtDNA mutations, which causes accelerated ageing104. In HGPS, mitochondria are aberrantly swollen and fragmented105, and OXPHOS activity progressively decreases from an early age106. Wrn-null mice have enlarged mitochondria, increased accumulation of mtDNA mutations and increased ROS formation at the expense of ATP generation107.

The OXPHOS system is a major producer of ROS108. Chronically elevated ROS levels lead to damage of cellular macromolecules, including OXPHOS complex proteins and mtDNA, creating a cycle in which ROS impairs mitochondrial integrity to further increase oxidative stress109 (FIG. 4). In Alzheimer disease, elevated ROS levels precede and promote the formation of β-amyloid- and tau-containing plaques and neurofibrillary tangles103, which further impair OXPHOS complexes110. A similar self-promoting loop occurs in familial Parkinson disease, in which disease-causing proteins increase ROS, impair OXPHOS and ultimately trigger neuronal death103. Genetic abrogation of the mtDNA surveillance factor TFAM similarly results in Parkinson disease in mice103. In atherosclerosis, chronic ROS promotes plaque formation by inhibiting OXPHOS and oxidizing low-density lipoprotein, which attracts monocytes and thereby has deleterious effects103. Mitochondrial dysfunction-induced inflammatory responses have also been reported in COPD111. ROS-induced loss of mitochondrial integrity further promotes vascular calcification, which is observed in individuals with HGPS, atherosclerosis, osteoporosis, CKD, T2D or Alzheimer disease112,113.

Decreased antioxidative mechanisms and mitochondrial biogenesis in HGPS and AADs.

Cells can disrupt deleterious ROS-mitochondrial interplay by activating antioxidative mechanisms that decrease ROS levels (FIG. 4). Nuclear factor erythroid 2-related factor 2 (NRF2) is a transcriptional master regulator of antioxidants, including thioredoxin-dependent peroxide reductase (PRDX3), which neutralizes most mitochondrially produced hydrogen peroxide114,115. NRF2 activity decreases with age116. Progerin aberrantly sequesters NRF2 and thereby inhibits its transcriptional activity, which results in chronic oxidative stress and drives the formation of cellular ageing in HGPS27. Defective antioxidative defence mechanisms also underlie Alzheimer disease and Parkinson disease, possibly either through defective NRF2 activation or via direct inhibition of antioxidants by disease-causing proteins, such as PRDX3 inhibition via a Parkinson disease-causing LRRK2 mutant108,115,117. Restoring mitochondrial integrity by overexpressing a mitochondria-targeted catalase antioxidant extends the lifespan of mice and protects against neurodegeneration, cardiac disease, cancer and insulin resistance103,118.

Increased mitochondrial turnover and replacement of damaged mitochondria with intact ones also protects against oxidative damage. Peroxisome proliferator-activated receptor-γ co-activator 1α (PGC1α) promotes mitochondrial biogenesis while simultaneously activating the NRF2 antioxidative response103 (FIG. 4). Activation of PGC1α alleviates mitochondrial dysfunction and other ageing defects in fibroblasts from patients with HGPS105, although the beneficial effects of PCG1α activation in vivo appear to be tissue-specific and dependent on the duration and extent of activation37. Decreased PGC1α activity occurs in Alzheimer disease and Parkinson disease and is implicated in CKD-associated muscular dystrophy110,119. Diminished metabolic activation of the PGC1α upstream regulators SIRT1 and AMPK in HGPS, ageing and AADs likely underlies the observed decrease in PGC1α activity26.

Overall, these observations indicate that metabolic signalling is coupled to mitochondrial integrity and ROS formation. The balance of this triad has major implications for the proteolytic autophagy system, which aids in the degradation of defective mitochondria, the recycling of amino acids required for mTOR-regulated protein synthesis and the removal of ROS-damaged macromolecules to maintain cellular homeostasis.

Proteostasis and proteolysis

Proteostasis is defined as the maintenance of a functional proteome through balanced regulation of protein synthesis, repair and proteolysis. The integrity of the proteome is dependent on continuous protein turnover, which ensures the degradation of damaged and misfolded proteins by autophagy and the ubiquitin–proteasome system (UPS), both of which are affected in human progeroid syndromes, physiological ageing and AADs.

Autophagy ameliorates ageing defects in HGPS

The metabolically driven build-up of damaged mitochondria and oxidized protein aggregates in ageing and disease indicates an imbalance between proteotoxic insults and the proteolytic pathways that remove damaged organelles and protein aggregates. Autophagy is a bulk degradative mechanism in which autophagosomes engulf damaged cellular components and promote their proteolytic degradation through fusion with hydrolase-containing lysosomes120.

Interestingly, the IIS cascade impairs autophagy by activating mTOR26 (FIG. 4). While temporary attenuation of autophagy by mTOR can be beneficial for cellular health by promoting anabolism, sustained autophagic impairment results in the toxic accumulation of protein aggregates in inclusion bodies called aggresomes, which fail to be degraded. Aggresomes contain both the autophagy adaptor protein p62 (also known as SQSTM1) and ubiquitin, which normally attract and promote proteolytic degradation of the aggregates120. In HGPS, ubiquitin- and p62-positive progerin aggregates have been identified121. Progerin aggregates are thought to be deleterious owing to the entrapment of normal cellular proteins, including NRF2 (REF. 27). Activation of autophagy through rapamycin-mediated mTOR inhibition promotes the solubilization and clearance of progerin and thereby alleviates cellular ageing defects121. Rapamycin treatment extends the lifespan of mice as well as that of C. elegans, where these lifespan-extending effects of IIS suppression are autophagy-dependent26,122.

Proteostasis and autophagy defects in AADs

There is evidence of benefits of autophagic activation in AADs, and these effects can mostly be attributed to the increased clearance of disease-promoting toxic aggregates that form during chronic oxidative stress120. In Alzheimer disease, the levels of the lysosomal pro-tease cathepsin D are reduced, impairing the clearance of β-amyloid and phosphorylated tau aggregates. Reducing the hyperactivated status of IIS–mTOR signalling increases autophagic clearance of these aggregates and rescues memory deficits in mouse models of Alzheimer disease123,124. Patients with Parkinson disease have decreased levels of lysosome-associated membrane glycoprotein 2 (LAMP2), which is thought to impair the clearance of α-synuclein aggregates and increase neuronal death125. In addition, genetic defects in the parkin and PTEN-induced putative kinase protein 1 (PINK1) proteins (BOX 1) impair autophagic clearance of damaged mitochondria in familial Parkinson disease120.

LAMP2A and cathepsin D deficiencies are also known causes of cardiomyopathies, and impaired autophagy can promote atherosclerosis by promoting VSMC senescence and cell death126,127. Protein aggregation has also been implicated in T2D, as it has been found that amylin aggregates form in the pancreas in T2D mouse models; furthermore, aggregation is increased by genetic ablation of the autophagic protein ATG7 (REF. 128). Furthermore, an amylin mutation that increases the propensity of the protein to form aggregates is associated with T2D. Mutation-induced amylin aggregation precedes pancreatic β-cell dysfunction and is sufficient to induce a diabetic phenotype in wild-type rats128.

Low proteasomal activity in ageing and AADs

The UPS clears damaged proteins through ATP-dependent proteolysis126. Patients with HGPS have decreased UPS activity29 (FIG. 4), and proteasomal activity declines with ageing. In line with these findings, increased UPS activity extends the lifespan of Drosophila melanogaster and C. elegans129.

Dopaminergic ablation of the P26S4 subunit of the UPS in mice increases α-synuclein aggregation and neuronal death in a manner similar to Parkinson disease120. In addition, reduced levels of the UPS activators PA700 and PA28 in the substantia nigra correlate with α-synuclein aggregation in patients with Parkinson disease130. Moderate repression of UPS activity by over expression of a dominant-negative proteasomal PSMB5 catalytic subunit mutant aggravates ischaemia–reperfusion-mediated cardiac defects, and pre-amyloid oligomer levels are increased in dilated and hyper-trophic cardiomyopathy126. Conversely, UPS activation is an emerging therapeutic strategy against cardiomyopathies126. Overall, these findings indicate that autophagy and UPS play a crucial part in counteracting ageing-associated metabolically-induced stress.

Perturbation of cell fate

The accumulation of cellular damage in ageing triggers pathways that control cell proliferation and differentiation as well as stem cell-mediated regeneration to prevent the permanent collapse of cellular and tissue homeostasis. Of particular importance in ageing and AADs are cellular senescence and regenerative pathways.

Increased cell senescence in HGPS and AADs

The loss of genomic and proteomic integrity can induce cellular senescence, which is an irreversible state of proliferative arrest that is mainly dependent on chronic activation of the tumour suppressor RB–p16 and p53–p21 pathways131,132 (FIG. 5). Temporary activation of these pathways can be beneficial, as it enables the repair of cellular damage before continuing proliferation. However, chronic activation of the RB–p16 and p53–p21 pathways has detrimental effects because it triggers self-sustaining inflammatory signalling driven by transforming growth factor β (TGFβ) and nuclear factor-κβ (NF-κB). This signalling is typified by the expression of interleukin-1 (IL-1) and 6 (IL-6), tumour necrosis factor-α (TNFα), cyclooxygenase 2 (COX2), and C-X-C motif chemokine 1 (CXCL1) and CXCL2 (REF. 131) (FIG. 5). This state is referred to as senescence-associated secretory phenotype (SASP), and its long-term activation can compromise the intercellular communication between senescent and immune-responsive cells, thereby preventing immune system-mediated clearance of senescent cells, which is normally promoted by SASP131. Failure to clear senescent cells can aggravate physiological decline during ageing by inducing cellular senescence in neighbouring cells via gap junction-mediated cell–cell contacts26. In line with this finding, artificially induced clearance of senescent cells extends the lifespan and healthspan of progeroid mice lacking the mitotic checkpoint protein BubR1 (TABLE 1), inhibits atherogenesis in low-density lipoprotein receptor-deficient mice and preserves renal and cardiac function in wild-type mice6,7,133.

Figure 5 |. Cellular senescence pathways.

Figure 5 |

Chronic levels of elevated DNA damage and metabolic-induced cellular stress trigger the activation of p16 and p53 in Hutchinson–Gilford progeria syndrome (HGPS), ageing and ageing-associated diseases (indicated by arrows), through which activation of RB and p21 results in permanent senescent growth arrest26. Increased activation of transforming growth factor β (TGFβ) and nuclear factor-κβ (NF-κB) drive the senescent-associated secretory phenotype (SASP), which is characterized by the indicated inflammatory regulators. COX2, cyclooxygenase 2; CXCL1, C-X-C motif chemokine 1; TNFα, tumour necrosis factor-α.

In HGPS, increased cellular damage hyperactivates the RB–p16, p53–p21 and NF-κB pathways, causing cellular senescence and SASP, which is accompanied by increased levels of IL6, TNFα, CXCL1 and CXCL2 in progeria mouse models34,134,135. NF-κB hyperactivation is partially detrimental owing to its epigenetics-altering effects via activation of the histone methyltransferase DOT1L136. Conversely, genetic disruption of NF-κB or p53 signalling improves the lifespan and healthspan of HGPS mouse models34,134.

In Alzheimer disease, neuron-supporting astrocytes, which have naturally high basal NF-κB activity and are implicated in the disease’s pathology, show increased p16 and IL-6 expression levels137. Hyperactivation of canonical NF-κB signalling is detrimental to astrocytes and is believed to promote β-amyloid aggregation, which reversibly drives TNFα-, IL-1β- and COX2-mediated inflammatory signalling138. Like Alzheimer disease, Parkinson disease leads to increased NF-κB activity and COX2 levels139. Treatment of wild-type mice with COX2 inhibitors attenuates the neurotoxic effects of MPTP (1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine), which induces Parkinson disease via the inhibition of OXPHOS139. Increased IL-1β exposure further aggravates the progression of Parkinson disease by inducing α-synuclein expression139.

In ageing-associated idiopathic lung fibrosis, increased IL-1β levels may promote the differentiation of lung fibroblasts into myofibroblasts, thereby promoting fibrosis140. SASP-regulated changes in cellular identity also play a part in cancer, as IL-6 and IL-8 promote tumorigenesis by increasing epithelial-to-mesenchymal transition (EMT)131. SASP-suppressing TGFβ inhibitors have been shown to attenuate the progression of some tumours and are currently being evaluated in clinical trials141. Finally, TGFβ- and NF-κB-driven inflammatory processes leading to SASP underlie CKD, and repression of both pathways with angiotensin-converting enzyme inhibitors slows the progression of CKD141.

Overall, irreversible growth arrest and perpetual hyperactivation of inflammatory secretory pathways can lead to a permanent state of cellular senescence that underlies ageing and AADs. The accumulation of senescent cells can be counteracted by stem cell-mediated regeneration.

Compromised stem cell regenerative capacity

Stem cells, which are capable of self-renewal and multi-lineage differentiation, reside in specific niches throughout the human body. Human mesenchymal stem cells (hMSCs) are present in virtually all tissues and differentiate into osteocytes, chondrocytes and adipocytes, whereas VSMCs are non-terminally differentiated cells of the vascular wall and only have limited differentiation plasticity142,143. Cellular ageing defects in fibroblasts isolated from patients with HGPS can be reversed upon generation of induced pluripotent stem cells (iPSCs), which do not express progerin because of the naturally occurring transcriptional repression of the LMNA gene that occurs in iPSCs30,31. Upon re-differentiation of HGPS-iPSCs into hMSCs and VSMCs, the transcriptional reactivation of the LMNA gene and the concomitant induction of progerin expression results in the emergence of the typical genomic, proteomic and senescent cellular ageing defects associated with HGPS30,31 (FIGS 14) and decreases survival rates under stress conditions, such as hypoxia. This relationship may explain the loss of VSMCs in patients with HGPS owing to exposure to haemodynamic stress as well as the reduced ability of HGPS-hMSCs to prevent ischaemia–reperfusion damage in a hindlimb muscle model31. The viability and capacity of HGPS-hMSCs to withstand chronic oxidative stress can be restored through re-activation of NRF2, which is impaired by progerin27.

In addition to stem cell exhaustion, which can also be observed for dermal stem cells in an HGPS mouse model, the accumulation of ageing defects may compromise the regenerative capacity of stem cells by altering their differentiation potential144. In that regard, progerin delays the differentiation of VSMCs and increases the osteogenic differentiation of hMSCs at the expense of adipocyte generation, which may contribute to the calcification of tissues (such as the skin and vascular walls) observed in patients with HGPS48,145.

The observed accumulation of low levels of progerin and decreased levels of Werner protein in wild-type VSMCs and hMSCs, respectively, from elderly individuals may also explain the impaired regenerative capacity of stem cells during physiological ageing82,146. Indeed, one or more of the typical cellular ageing hallmarks, such as oxidative stress or activation of DNA damage, senescence and SASP pathways, has been reported to occur with ageing in hMSCs, VSMCs, haematopoietic stem cells (HSCs), skeletal muscle cells and neuronal stem cells143,147. The detrimental effects of these hallmarks on stem cells is demonstrated by the decreased regrowth of lung and pancreatic tissues, which have naturally high regenerative potential, upon partial resection in aged humans compared with that in young adults148,149. Conversely, silencing p16, activating SIRT1, blocking SASP effects by overexpressing an IL-1 receptor antagonist or reducing ROS by overexpressing the anti-oxidant superoxide dismutase 2 improves the regenerative capacity of hMSCs and HSCs as well as neuronal, intestinal and muscle stem cells143,147.

Adult stem cells may be particularly prone to ageing defects, as they mostly exist in quiescence, that is, a state of non-permanent growth arrest, which has been reported to coincide with decreased levels of DNA repair proteins and the postponement of DNA repair until re-entry into the cell cycle147. It is conceivable that the accumulation of DNA damage during this quiescent state may exceed a threshold that impairs subsequent stem cell function.

Decreased regenerative capacity of stem cells is important in several AADs. For example, VSMC senescence promotes the formation of atherosclerotic plaques150. In T2D, hMSCs may have diminished proliferative and differentiation capacity, and hMSC- and pancreatic-specific stem cell activity is beneficial to pancreatic β-cell mass and function148. Furthermore, the differentiation potential of lung-specific stem cells is dysregulated in ageing-associated idiopathic pulmonary fibrosis, and functional hMSCs aid in the repair of osteoporotic bone fractures143,149. Although the contribution of cardiac and neuronal stem cells to the turnover of terminally differentiated myocytes and neurons remains unclear147,151, their beneficial immunoregulatory secretory effects have been established and shown to alleviate cognitive defects in mouse models of Alzheimer disease upon injection of hMSCs147,151.

Overall, these observations in premature ageing disorders, natural ageing and AADs suggest that premature ageing and AADs are driven by genetic, metabolic and senescence-associated cellular-ageing defects that perturb the fate of stem cells.

Perspective

Here, we have explored the interplay between premature ageing, normal ageing and AADs. We discuss how individual cellular defects in the regulation of genomic, telomeric, epigenetic, metabolic, mitochondrial, proteolytic and cell cycle integrity contribute to cellular ageing and AADs (FIG. 1). The picture that emerges is one of extensive connections between these hallmarks of ageing; consequently, disrupting one of these processes induces other defects, which ultimately lead to a collapse of cell homeostasis that drives disease. Although this cascade reduces healthspan and drives ageing, the interconnected nature of these events may also provide a therapeutic opportunity in that amelioration of one or more of these hallmarks of ageing may be sufficient to alleviate the process as a whole.

Observations made in studies of premature ageing systems, normal ageing and AADs are complementary to each other. Premature ageing diseases only reflect some aspects of ageing, but they offer the benefit of well-defined genetics and controllable experimental approaches. Physiological ageing and AADs reflect the true ageing process, but their study is complicated by the confounding effects of their physiological and environmental context. Further unravelling the intricacies of the cellular ageing network will require the development of novel experimental model systems that are not hampered by the traditionally slow stochastic accumulation and low penetrance of ageing defects. Premature ageing diseases in humans provide such an opportunity because the associated cellular ageing defects recapitulate, to a large extent, those observed in physiological ageing and AADs. Furthermore, the accelerated nature of premature ageing syndromes, their defined single genetic cause and their strong phenotypic penetrance enables the development of robust cellular ageing assays.

The generation of iPSCs from patients with these and other premature ageing diseases is an exciting new strategy that enables the temporal dissection of the formation of prominent ageing defects152. Alternatively, inducible expression of mutant proteins linked to progeroid syndromes in wild-type cells can similarly address these questions, and both strategies have the additional benefit of being easily combined with available RNAi and CRISPR high-throughput screening approaches. CRISPR-Cas9 gene editing can further accelerate the characterization of newly identified ageing mechanisms by facilitating the generation of novel in vivo ageing and AAD models.

We anticipate that the development of new experimental approaches for the study of human premature ageing diseases will begin to resolve key questions on how cellular ageing defects are linked and how they are counteracted by anti-ageing pathways. Such studies will increase our understanding of the ageing process and will facilitate the detection and treatment of chronic ageing-associated and leading diseases, such as Alzheimer disease, Parkinson disease, atherosclerosis, diabetes and cancer.

Glossary

Nucleotide excision repair

(NER). A DNA repair pathway specialized in the removal of bulky DNA adducts, including ultraviolet damage-induced thymidine dimers

Non-homologous end joining

(NHEJ). A DNA repair pathway that repairs double-strand breaks through direct ligation of the broken ends without the need of a homologous template

Oxidative phosphorylation

(OXPHOS). A process by which electrons are transferred from electron donors to electron acceptors, thereby releasing energy in the form of ATP. In prokaryotes, this process takes place in the inner mitochondrial membrane at the site of the electron transport chain

Ubiquitin–proteasome system

(UPS). A system that degrades proteins marked by degradation-specific ubiquitin marks in an ATP-dependent manner

Epithelial-to-mesenchymal transition

(EMT). A process by which epithelial cells undergo various molecular changes related to cell–cell adhesion, polarity and invasive properties, in order to become mesenchymal cells. EMT has beneficial roles in wound healing but exerts detrimental effects in organ fibrosis and the initiation of tumour metastasis

Footnotes

Competing interests statement

The authors declare no competing interests.

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

References

  • 1.Lopez AD, Mathers CD, Ezzati M, Jamison DT & Murray CJ Global and regional burden of disease and risk factors, 2001: systematic analysis of population health data. Lancet 367, 1747–1757 (2006). [DOI] [PubMed] [Google Scholar]
  • 2.National Center for Health Statistics. Deaths and mortality. Center for Disease Control and Prevention http://www.cdc.gov/nchs/fastats/deaths.htm (2017).
  • 3.Denayer T, Stöhr T & Van Roy M Animal models in translational medicine: validation and prediction. New Horizons Transl Med. 2, 5–11 (2014). [Google Scholar]
  • 4.Potashkin JA, Blume SR & Runkle NK Limitations of animal models of Parkinson’s disease. Parkinsons Dis. 2011, 658083 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Getz GS & Reardon CA Animal models of atherosclerosis. Arterioscler. Thromb. Vasc. Biol 32, 1104–1115 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Childs BG et al. Senescent intimal foam cells are deleterious at all stages of atherosclerosis. Science 354, 472–477 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Baker DJ et al. Naturally occurring p16Ink4a-positive cells shorten healthy lifespan. Nature 530, 184–189 (2016).An important study elegantly demonstrating the detrimental effect of the accumulation of senescent cells on lifespan and healthspan.
  • 8.Nystrom H, Nordstrom A & Nordstrom P Risk of injurious fall and hip fracture up to 26 y before the diagnosis of Parkinson disease: nested case-control studies in a nationwide cohort. PLoS Med. 13, e1001954 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Chang YT et al. Handgrip strength is an independent predictor of renal outcomes in patients with chronic kidney diseases. Nephrol. Dial. Transplant 26, 3588–3595 (2011). [DOI] [PubMed] [Google Scholar]
  • 10.Eriksson M et al. Recurrent de novo point mutations in lamin A cause Hutchinson–Gilford progeria syndrome. Nature 423, 293–298 (2003). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.de Sandre-Giovannoli A et al. Lamin A truncation in Hutchinson–Gilford progeria. Science 1084, 2055 (2003). [DOI] [PubMed] [Google Scholar]
  • 12.Yu CE et al. Positional cloning of the Werner’s syndrome gene. Science 272, 258–262 (1996). [DOI] [PubMed] [Google Scholar]
  • 13.Hennekam RC Hutchinson–Gilford progeria syndrome: review of the phenotype. Am. J. Med. Genet. A 140, 2603–2624 (2006). [DOI] [PubMed] [Google Scholar]
  • 14.Gordon LB et al. Impact of farnesylation inhibitors on survival in Hutchinson–Gilford progeria syndrome. Circulation 130, 27–34 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Gordon LB, Rothman FG, Lopez-Otin C & Misteli T Progeria: a paradigm for translational medicine. Cell 156, 400–407 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Dittmer TA & Misteli T The lamin protein family. Genome Biol. 12, 222 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Davies BS, Fong LG, Yang SH, Coffinier C & Young SG The posttranslational processing of prelamin A and disease. Annu. Rev. Genomics Hum. Genet 10, 153–174 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Taimen P et al. A progeria mutation reveals functions for lamin A in nuclear assembly, architecture, and chromosome organization. Proc. Natl Acad. Sci. USA 106, 20788–20793 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Dahl KN, Ribeiro AJ & Lammerding J Nuclear shape, mechanics, and mechanotransduction. Circ. Res 102, 1307–1318 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Carrero D & Soria-Valles C Hallmarks of progeroid syndromes: lessons from mice and reprogrammed cells. Dis. Model. Mech 9, 719–735 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.de la Rosa J et al. Prelamin A causes progeria through cell-extrinsic mechanisms and prevents cancer invasion. Nat. Commun 4, 2268 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Jung HJ et al. Regulation of prelamin A but not lamin C by miR-9, a brain-specific microRNA. Proc. Natl Acad. Sci. USA 109, E423–E431 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Fernandez P et al. Transformation resistance in a premature aging disorder identifies a tumor-protective function of BRD4. Cell Rep. 9, 248–260 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Merideth MA et al. Phenotype and course of Hutchinson–Gilford progeria syndrome. N. Engl. J. Med 358, 592–604 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Broers JL, Ramaekers FC, Bonne G, Yaou RB & Hutchison CJ Nuclear lamins: laminopathies and their role in premature ageing. Physiol. Rev 86, 967–1008 (2006). [DOI] [PubMed] [Google Scholar]
  • 26.Lopez-Otin C, Blasco MA, Partridge L, Serrano M & Kroemer G The hallmarks of aging. Cell 153, 1194–1217 (2013).A comprehensive overview of the cellular defects that occur during ageing.
  • 27.Kubben N et al. Repression of the antioxidant Nrf2 pathway in premature aging. Cell 165, 1361–1374 (2016).An important study that used unbiased screening to identify the mechanisms that drive premature cellular ageing. Entrapment of the antioxidant-promoting transcription factor NRF2 by progerin was found as a novel mechanism that drives HGPS and is likely a cause of chronic oxidative stress during ageing.
  • 28.Scaffidi P & Misteli T Lamin A-dependent nuclear defects in human aging. Science 312, 1059–1063 (2006).This study provided the first evidence that similar changes in histone modifications occur in HGPS and normal ageing.
  • 29.Viteri G, Chung YW & Stadtman ER Effect of progerin on the accumulation of oxidized proteins in fibroblasts from Hutchinson Gilford progeria patients. Mech. Ageing Dev 131, 2–8 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Liu GH et al. Recapitulation of premature ageing with iPSCs from Hutchinson–Gilford progeria syndrome. Nature 472, 221–225 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Zhang J et al. A human iPSC model of Hutchinson Gilford Progeria reveals vascular smooth muscle and mesenchymal stem cell defects. Cell Stem Cell 8, 31–45 (2011). [DOI] [PubMed] [Google Scholar]
  • 32.Pegoraro G et al. Ageing-related chromatin defects through loss of the NURD complex. Nat. Cell Biol 11, 1261–1267 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Van Berlo JH et al. A-Type lamins are essential for TGF-beta1 induced PP2A to dephosphorylate transcription factors. Hum. Mol. Genet 14, 2839–2849 (2005). [DOI] [PubMed] [Google Scholar]
  • 34.Varela I et al. Accelerated ageing in mice deficient in Zmpste24 protease is linked to p53 signalling activation. Nature 437, 564–568 (2005).This study provided important proof that chronic activation of the p53 pathway as a response to elevated cellular stress has a detrimental effect in prelamin A-induced premature ageing.
  • 35.Puente XS et al. Exome sequencing and functional analysis identifies BANF1 mutation as the cause of a hereditary progeroid syndrome. Am. J. Hum. Genet 88, 650–656 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.McClintock D et al. The mutant form of lamin A that causes Hutchinson–Gilford progeria is a biomarker of cellular aging in human skin. PLoS ONE 2, e1269 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Lopez-Mejia IC et al. Antagonistic functions of LMNA isoforms in energy expenditure and lifespan. EMBO Rep. 15, 529–539 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Hisama FM et al. Coronary artery disease in a Werner syndrome-like form of progeria characterized by low levels of progerin, a splice variant of lamin A. Am J Med Genet A 155, 3002–3006 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Atamna H, Cheung I & Ames BN A method for detecting abasic sites in living cells: age-dependent changes in base excision repair. Proc. Natl Acad. Sci. USA 97, 686–691 (2000). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Annett K et al. An investigation of DNA mismatch repair capacity under normal culture conditions and under conditions of supra-physiological challenge in human CD4+T cell clones from donors of different ages. Exp. Gerontol 40, 976–981 (2005). [DOI] [PubMed] [Google Scholar]
  • 41.Moriwaki S, Ray S, Tarone RE, Kraemer KH & Grossman L The effect of donor age on the processing of UV-damaged DNA by cultured human cells: reduced DNA repair capacity and increased DNA mutability. Mutat. Res 364, 117–123 (1996). [DOI] [PubMed] [Google Scholar]
  • 42.Mao Z et al. Sirtuin 6 (SIRT6) rescues the decline of homologous recombination repair during replicative senescence. Proc. Natl Acad. Sci. USA 109, 11800–11805 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Stuart GR & Glickman BW Through a glass, darkly: reflections of mutation from lacI transgenic mice. Genetics 155, 1359–1367 (2000). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Liu Y et al. Involvement of xeroderma pigmentosum group A (XPA) in progeria arising from defective maturation of prelamin A. FASEB J. 22, 603–611 (2008).This paper provides a mechanistic explanation for the frequently observed persistence of irreparable DNA damage in HGPS.
  • 45.Musich PR & Zou Y Genomic instability and DNA damage responses in progeria arising from defective maturation of prelamin A. Aging (Albany NY) 1, 28–37 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Tang H, Hilton B, Musich PR, Fang DZ & Zou Y Replication factor C1, the large subunit of replication factor C, is proteolytically truncated in Hutchinson–Gilford progeria syndrome. Aging Cell 11, 363–365 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Cobb AM, Murray TV, Warren DT, Liu Y & Shanahan CM Disruption of PCNA-lamins A/C interactions by prelamin A induces DNA replication fork stalling. Nucleus 7, 498–511 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Zhang H, Xiong ZM & Cao K Mechanisms controlling the smooth muscle cell death in progeria via down-regulation of poly(ADP-ribose) polymerase 1. Proc. Natl Acad. Sci. USA 111, E2261–E2270 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Liu B et al. Genomic instability in laminopathy-based premature aging. Nat. Med 11, 780–785 (2005). [DOI] [PubMed] [Google Scholar]
  • 50.Rass U, Ahel I & West SC Defective DNA repair and neurodegenerative disease. Cell 130, 991–1004 (2007). [DOI] [PubMed] [Google Scholar]
  • 51.Shackelford DA DNA end joining activity is reduced in Alzheimer’s disease. Neurobiol. Aging 27, 596–605 (2006). [DOI] [PubMed] [Google Scholar]
  • 52.Kao SY Regulation of DNA repair by parkin. Biochem. Biophys. Res. Commun 382, 321–325 (2009). [DOI] [PubMed] [Google Scholar]
  • 53.Robbins JH et al. Parkinson’s disease and Alzheimer’s disease: hypersensitivity to X rays in cultured cell lines. J. Neurol. Neurosurg. Psychiatry 48, 916–923 (1985). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Mahmoudi M, Mercer J & Bennett M DNA damage and repair in atherosclerosis. Cardiovasc. Res 71, 259–268 (2006). [DOI] [PubMed] [Google Scholar]
  • 55.Muftuoglu M et al. The clinical characteristics of Werner syndrome: molecular and biochemical diagnosis. Hum. Genet 124, 369–377 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Tavana O et al. Ku70 functions in addition to nonhomologous end joining in pancreatic beta-cells: a connection to beta-catenin regulation. Diabetes 62, 2429–2438 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Schmidt E et al. Expression of the Hutchinson–Gilford progeria mutation during osteoblast development results in loss of osteocytes, irregular mineralization, and poor biomechanical properties. J. Biol. Chem 287, 33512–33522 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Lombard DB et al. DNA repair, genome stability, and aging. Cell 120, 497–512 (2005). [DOI] [PubMed] [Google Scholar]
  • 59.Gorbunova V, Seluanov A, Mao Z & Hine C Changes in DNA repair during aging. Nucleic Acids Res. 35, 7466–7474 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Fuster JJ & Andres V Telomere biology and cardiovascular disease. Circ. Res 99, 1167–1180 (2006). [DOI] [PubMed] [Google Scholar]
  • 61.Wong LS et al. Telomere biology in cardiovascular disease: the TERC−/− mouse as a model for heart failure and ageing. Cardiovasc. Res 81, 244–252 (2009). [DOI] [PubMed] [Google Scholar]
  • 62.Benson EK, Lee SW & Aaronson SA Role of progerin-induced telomere dysfunction in HGPS premature cellular senescence. J. Cell Sci 123, 2605–2612 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Decker ML, Chavez E, Vulto I & Lansdorp PM Telomere length in Hutchinson–Gilford progeria syndrome. Mech. Ageing Dev 130, 377–383 (2009). [DOI] [PubMed] [Google Scholar]
  • 64.Gonzalez-Suarez I et al. Novel roles for A-type lamins in telomere biology and the DNA damage response pathway. EMBO J. 28, 2414–2427 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.De Vos WH et al. Increased plasticity of the nuclear envelope and hypermobility of telomeres due to the loss of A-type lamins. Biochim. Biophys. Acta 1800, 448–458 (2010). [DOI] [PubMed] [Google Scholar]
  • 66.Raz V et al. The nuclear lamina promotes telomere aggregation and centromere peripheral localization during senescence of human mesenchymal stem cells. J. Cell Sci 121, 4018–4028 (2008). [DOI] [PubMed] [Google Scholar]
  • 67.Wood AM et al. TRF2 and lamin A/C interact to facilitate the functional organization of chromosome ends. Nat. Commun 5, 5467 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Saha B et al. DNA damage accumulation and TRF2 degradation in atypical Werner syndrome fibroblasts with LMNA mutations. Front. Genet 4, 129 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Cao K et al. Progerin and telomere dysfunction collaborate to trigger cellular senescence in normal human fibroblasts. J. Clin. Invest 121, 2833–2844 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Li D, Yuan Q & Wang W The role of telomeres in musculoskeletal diseases. J. Int. Med. Res 40, 1242–1250 (2012). [DOI] [PubMed] [Google Scholar]
  • 71.Kordinas V, Ioannidis A & Chatzipanagiotou S The telomere/telomerase system in chronic inflammatory diseases. Cause or effect? Genes (Basel) 7, 60 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Alder JK et al. Short telomeres are a risk factor for idiopathic pulmonary fibrosis. Proc. Natl Acad. Sci. USA 105, 13051–13056 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Adnot S et al. Telomere dysfunction and cell senescence in chronic lung diseases: therapeutic potential. Pharmacol. Ther 153, 125–134 (2015). [DOI] [PubMed] [Google Scholar]
  • 74.Chojnowski A et al. Progerin reduces LAP2alpha-telomere association in Hutchinson–Gilford progeria. eLife 4, e07759 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Heyn H, Moran S & Esteller M Aberrant DNA methylation profiles in the premature aging disorders Hutchinson–Gilford Progeria and Werner syndrome. Epigenetics 8, 28–33 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Scaffidi P & Misteli T Reversal of the cellular phenotype in the premature aging disease Hutchinson–Gilford progeria syndrome. Nat. Med 11, 440–445 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Shumaker DK et al. Mutant nuclear lamin A leads to progressive alterations of epigenetic control in premature aging. Proc. Natl Acad. Sci. USA 103, 8703–8708 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Jin C et al. Histone demethylase UTX-1 regulates C. elegans life span by targeting the insulin/IGF-1 signaling pathway. Cell Metab. 14, 161–172 (2011). [DOI] [PubMed] [Google Scholar]
  • 79.Chen H et al. Polycomb protein Ezh2 regulates pancreatic beta-cell Ink4a/Arf expression and regeneration in diabetes mellitus. Genes Dev. 23, 975–985 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Lund HL & van Loohuizen M Polycomb complexes and silencing mechanisms. Curr. Opin. Cell Biol 16, 239–246 (2004). [DOI] [PubMed] [Google Scholar]
  • 81.Marullo F et al. Nucleoplasmic Lamin A/C and Polycomb group of proteins: an evolutionarily conserved interplay. Nucleus 7, 103–111 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Zhang W et al. A Werner syndrome stem cell model unveils heterochromatin alterations as a driver of human aging. Science 348, 1160–1163 (2015).This paper demonstrates that global epigenetic defects underlie mesenchymal stem cell dysfunction in Werner premature ageing syndrome.
  • 83.Liu B et al. Depleting the methyltransferase Suv39h1 improves DNA repair and extends lifespan in a progeria mouse model. Nat. Commun 4, 1868 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Wing MR, Ramezani A, Gill HS, Devaney JM & Raj DS Epigenetics of progression of chronic kidney disease: fact or fantasy? Semin. Nephrol 33, 363–374 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Frost B, Hemberg M, Lewis J & Feany MB Tau promotes neurodegeneration through global chromatin relaxation. Nat. Neurosci 17, 357–366 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Misteli T Higher-order genome organization in human disease. Cold Spring Harb. Perspect. Biol 2, a000794 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Krishnan V et al. Histone H4 lysine 16 hypoacetylation is associated with defective DNA repair and premature senescence in Zmpste24-deficient mice. Proc. Natl Acad. Sci. USA 108, 12325–12330 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Sharma GG et al. MOF and histone H4 acetylation at lysine 16 are critical for DNA damage response and double-strand break repair. Mol. Cell. Biol 30, 3582–3595 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Dang W et al. Histone H4 lysine 16 acetylation regulates cellular lifespan. Nature 459, 802–807 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Peleg S et al. Altered histone acetylation is associated with age-dependent memory impairment in mice. Science 328, 753–756 (2010). [DOI] [PubMed] [Google Scholar]
  • 91.Qiu X, Xiao X, Li N & Li Y Histone deacetylases inhibitors (HDACis) as novel therapeutic application in various clinical diseases. Prog. Neuropsychopharmacol. Biol. Psychiatry 72, 60–72 (2017). [DOI] [PubMed] [Google Scholar]
  • 92.Mazucanti CH et al. Longevity pathways (mTOR, SIRT, Insulin/IGF-1) as key modulatory targets on aging and neurodegeneration. Curr. Top. Med. Chem 15, 2116–2138 (2015). [DOI] [PubMed] [Google Scholar]
  • 93.Mariño G et al. Insulin-like growth factor 1 treatment extends longevity in a mouse model of human premature aging by restoring somatotroph axis function. Proc. Natl Acad. Sci. USA 107, 16268–16273 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Perluigi M, Di Domenico F & Butterfield DA mTOR signaling in aging and neurodegeneration: at the crossroad between metabolism dysfunction and impairment of autophagy. Neurobiol. Dis 84, 39–49 (2015). [DOI] [PubMed] [Google Scholar]
  • 95.Talaei F, van Praag VM & Henning RH Hydrogen sulfide restores a normal morphological phenotype in Werner syndrome fibroblasts, attenuates oxidative damage and modulates mTOR pathway. Pharmacol. Res 74, 34–44 (2013). [DOI] [PubMed] [Google Scholar]
  • 96.Sharples AP et al. Longevity and skeletal muscle mass: the role of IGF signalling, the sirtuins, dietary restriction and protein intake. Aging Cell 14, 511–523 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Massudi H et al. Age-associated changes in oxidative stress and NAD+ metabolism in human tissue. PLoS ONE 7, e42357 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Liu B et al. Resveratrol rescues SIRT1-dependent adult stem cell decline and alleviates progeroid features in laminopathy-based progeria. Cell Metab. 16, 738–750 (2012). [DOI] [PubMed] [Google Scholar]
  • 99.Jeon SM Regulation and function of AMPK in physiology and diseases. Exp. Mol. Med 48, e245 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Satriano J, Sharma K, Blantz RC & Deng A Induction of AMPK activity corrects early pathophysiological alterations in the subtotal nephrectomy model of chronic kidney disease. Am. J. Physiol. Renal Physiol 305, F727–F733 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Anderson JG et al. Enhanced insulin sensitivity in skeletal muscle and liver by physiological overexpression of SIRT6. Mol. Metab 4, 846–856 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Endisha H et al. Restoring SIRT6 expression in Hutchinson–Gilford progeria syndrome cells impedes premature senescence and formation of dysmorphic nuclei. Pathobiology 82, 9–20 (2015). [DOI] [PubMed] [Google Scholar]
  • 103.Lane RK, Hilsabeck T & Rea SL The role of mitochondrial dysfunction in age-related diseases. Biochim. Biophys. Acta 1847, 1387–1400 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Trifunovic A et al. Premature ageing in mice expressing defective mitochondrial DNA polymerase. Nature 429, 417–423 (2004).This study provided direct evidence that the accumulation of mutations in mtDNA that occurs with ageing is not simply correlative but causesa eduction in lifespan and leads to the development of premature ageing pathologies.
  • 105.Xiong ZM et al. Methylene blue alleviates nuclear and mitochondrial abnormalities in progeria. Aging Cell 15, 279–290 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Rivera-Torres J et al. Identification of mitochondrial dysfunction in Hutchinson–Gilford progeria syndrome through use of stable isotope labeling with amino acids in cell culture. J. Proteomics 91, 466–477 (2013). [DOI] [PubMed] [Google Scholar]
  • 107.Massip L et al. Vitamin C restores healthy aging in a mouse model for Werner syndrome. FASEB J. 24, 158–172 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Jiang T, Sun Q & Chen S Oxidative stress: a major pathogenesis and potential therapeutic target of antioxidative agents in Parkinson’s disease and Alzheimer’s disease. Prog. Neurobiol 147, 1–19 (2016). [DOI] [PubMed] [Google Scholar]
  • 109.de Souza-Pinto NC, Wilson DM III, Stevnsner TV & Bohr VA Mitochondrial DNA, base excision repair and neurodegeneration. DNA Repair (Amst.) 7, 1098–1109 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Golpich M et al. Mitochondrial dysfunction and biogenesis in neurodegenerative diseases: pathogenesis and treatment. CNS Neurosci. Ther 23, 5–22 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Kang MJ & Shadel GS A mitochondrial perspective of chronic obstructive pulmonary disease pathogenesis. Tuberc. Respir. Dis. (Seoul) 79, 207–213 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Villa-Bellosta R et al. Defective extracellular pyrophosphate metabolism promotes vascular calcification in a mouse model of Hutchinson–Gilford progeria syndrome that is ameliorated on pyrophosphate treatment. Circulation 127, 2442–2451 (2013). [DOI] [PubMed] [Google Scholar]
  • 113.Demer LL & Tintut Y Vascular calcification: pathobiology of a multifaceted disease. Circulation 117, 2938–2948 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Dinkova-Kostova AT & Abramov AY The emerging role of Nrf2 in mitochondrial function. Free Radic. Biol. Med 88, 179–188 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Pallardo FV et al. Mitochondrial dysfunction in some oxidative stress-related genetic diseases: Ataxia–Telangiectasia, Down Syndrome, Fanconi Anaemia and Werner Syndrome. Biogerontology 11, 401–419 (2010). [DOI] [PubMed] [Google Scholar]
  • 116.Suh JH et al. Decline in transcriptional activity of Nrf2 causes age-related loss of glutathione synthesis, which is reversible with lipoic acid. Proc. Natl Acad. Sci. USA 101, 3381–3386 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Ramsey CP et al. Expression of Nrf2 in neurodegenerative diseases. J. Neuropathol. Exp. Neurol 66, 75–85 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Schriner SE et al. Extension of murine life span by overexpression of catalase targeted to mitochondria. Science 308, 1909–1911 (2005).This paper provides evidence that mitochondria are a prominent source of oxidative stress that drives ageing.
  • 119.Tamaki M et al. Chronic kidney disease reduces muscle mitochondria and exercise endurance and its exacerbation by dietary protein through inactivation of pyruvate dehydrogenase. Kidney Int. 85, 1330–1339 (2014). [DOI] [PubMed] [Google Scholar]
  • 120.Tanaka K & Matsuda N Proteostasis and neurodegeneration: the roles of proteasomal degradation and autophagy. Biochim. Biophys. Acta 1843, 197–204 (2014). [DOI] [PubMed] [Google Scholar]
  • 121.Cao K et al. Rapamycin reverses cellular phenotypes and enhances mutant protein clearance in Hutchinson–Gilford progeria syndrome cells. Sci. Transl Med 3, 89ra58 (2011).This paper provides a mechanistic link between metabolic signalling and ageing via the autophagic degradation of progerin and offers a novel therapeutic strategy for HGPS that is currently being tested in a clinical trial.
  • 122.Anisimov VN et al. Rapamycin increases lifespan and inhibits spontaneous tumorigenesis in inbred female mice. Cell Cycle 10, 4230–4236 (2011). [DOI] [PubMed] [Google Scholar]
  • 123.Di Domenico F, Tramutola A & Perluigi M Cathepsin D as a therapeutic target in Alzheimer’s disease. Expert Opin. Ther. Targets 20, 1393–1395 (2016). [DOI] [PubMed] [Google Scholar]
  • 124.Caccamo A, De Pinto V, Messina A, Branca C & Oddo S Genetic reduction of mammalian target of rapamycin ameliorates Alzheimer’s disease-like cognitive and pathological deficits by restoring hippocampal gene expression signature. J. Neurosci 34, 7988–7998 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Xilouri M et al. Impairment of chaperone-mediated autophagy induces dopaminergic neurodegeneration in rats. Autophagy 12, 2230–2247 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Wang X & Robbins J Proteasomal and lysosomal protein degradation and heart disease. J. Mol. Cell. Cardiol 71, 16–24 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Tai S, Hu XQ, Peng DQ, Zhou SH & Zheng XL The roles of autophagy in vascular smooth muscle cells. Int. J. Cardiol 211, 1–6 (2016). [DOI] [PubMed] [Google Scholar]
  • 128.Mukherjee A, Morales-Scheihing D, Butler PC & Soto C Type 2 diabetes as a protein misfolding disease. Trends Mol. Med 21, 439–449 (2015).An interesting perspective on T2D, which has joined the growing ranks of diseases related to protein misfolding.
  • 129.Saez I & Vilchez D The mechanistic links between proteasome activity, aging and age-related diseases. Curr. Genomics 15, 38–51 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Deger JM, Gerson JE & Kayed R The interrelationship of proteasome impairment and oligomeric intermediates in neurodegeneration. Aging Cell 14, 715–724 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Tchkonia T, Zhu Y, van Deursen J, Campisi J & Kirkland JL Cellular senescence and the senescent secretory phenotype: therapeutic opportunities. J. Clin. Invest 123, 966–972 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Krishnamurthy J et al. Ink4a/Arf expression is a biomarker of aging. J. Clin. Invest 114, 1299–1307 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Baker DJ et al. Clearance of p16Ink4a-positive senescent cells delays ageing-associated disorders. Nature 479, 232–236 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Osorio FG et al. Nuclear lamina defects cause ATM-dependent NF-kappaB activation and link accelerated aging to a systemic inflammatory response. Genes Dev. 26, 2311–2324 (2012).An important study depicting for the first time that chronic activation of inflammatory pathways is crucial for the formation of premature ageing defects in HGPS.
  • 135.Ibrahim MX et al. Targeting isoprenylcysteine methylation ameliorates disease in a mouse model of progeria. Science 340, 1330–1333 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Soria-Valles C et al. NF-kappaB activation impairs somatic cell reprogramming in ageing. Nat. Cell Biol 17, 1004–1013 (2015). [DOI] [PubMed] [Google Scholar]
  • 137.Bhat R et al. Astrocyte senescence as a component of Alzheimer’s disease. PLoS ONE 7, e45069 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Shi ZM et al. Upstream regulators and downstream effectors of NF-kappaB in Alzheimer’s disease. J. Neurol. Sci 366, 127–134 (2016). [DOI] [PubMed] [Google Scholar]
  • 139.Phani S, Loike JD & Przedborski S Neurodegeneration and inflammation in Parkinson’s disease. Parkinsonism Relat. Disord 18 (Suppl. 1), S207–S209 (2012). [DOI] [PubMed] [Google Scholar]
  • 140.Kuwano K et al. Cellular senescence and autophagy in the pathogenesis of chronic obstructive pulmonary disease (COPD) and idiopathic pulmonary fibrosis (IPF). Respir. Investig 54, 397–406 (2016). [DOI] [PubMed] [Google Scholar]
  • 141.Decleves AE & Sharma K Novel targets of antifibrotic and anti-inflammatory treatment in CKD. Nat. Rev. Nephrol 10, 257–267 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Iyemere VP, Proudfoot D, Weissberg PL & Shanahan CM Vascular smooth muscle cell phenotypic plasticity and the regulation of vascular calcification. J. Intern. Med 260, 192–210 (2006). [DOI] [PubMed] [Google Scholar]
  • 143.Turinetto V, Vitale E & Giachino C Senescence in human mesenchymal stem cells: functional changes and implications in stem cell-based therapy. Int. J. Mol. Sci 17, E1164 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Rosengardten Y, McKenna T, Grochova D & Eriksson M Stem cell depletion in Hutchinson–Gilford progeria syndrome. Aging Cell 10, 1011–1020 (2011). [DOI] [PubMed] [Google Scholar]
  • 145.Scaffidi P & Misteli T Lamin A-dependent misregulation of adult stem cells associated with accelerated ageing. Nat. Cell Biol 10, 452–459 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Olive M et al. Cardiovascular pathology in Hutchinson–Gilford progeria: correlation with the vascular pathology of aging. Arterioscler. Thromb. Vasc. Biol 30, 2301–2309 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Oh J, Lee YD & Wagers AJ Stem cell aging: mechanisms, regulators and therapeutic opportunities. Nat. Med 20, 870–880 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Bhartiya D & Patel H Very small embryonic-like stem cells are involved in pancreatic regeneration and their dysfunction with age may lead to diabetes and cancer. Stem Cell Res. Ther 6, 96 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Navarro S & Driscoll B Regeneration of the aging lung: a mini-review. Gerontology 63, 270–280 (2016). [DOI] [PubMed] [Google Scholar]
  • 150.Wang J et al. Vascular smooth muscle cell senescence promotes atherosclerosis and features of plaque vulnerability. Circulation 132, 1909–1919 (2015). [DOI] [PubMed] [Google Scholar]
  • 151.Tincer G, Mashkaryan V, Bhattarai P & Kizil C Neural stem/progenitor cells in Alzheimer’s disease. Yale J. Biol. Med 89, 23–35 (2016). [PMC free article] [PubMed] [Google Scholar]
  • 152.Ocampo A et al. In vivo amelioration of age-associated hallmarks by partial reprogramming. Cell 167, 1719–1733.e12 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES