You are on page 1of 12

Endocrinology, 2021, Vol. 162, No.

10, 1–12
https://doi.org/10.1210/endocr/bqab136
Mini-Review

Mini-Review

Role of Cellular Senescence in Type II Diabetes


Akilavalli Narasimhan,1,* Rafael R. Flores,1,* Paul D. Robbins,1 and
Laura J. Niedernhofer1

Downloaded from https://academic.oup.com/endo/article/162/10/bqab136/6345039 by guest on 24 March 2024


1
Institute on the Biology of Aging and Metabolism, Department of Biochemistry, Molecular Biology, and
Biophysics, University of Minnesota Medical School, 55455, USA
ORCiD numbers: 0000-0003-1068-7099 (P. D. Robbins); 0000-0002-5580-6381 (L. J. Niedernhofer).

*These authors contributed equally

Abbreviations: AGE, advanced glycation end product; AT, adipose tissue; BAT, brown adipose tissue; DDR, DNA damage
response; DR, diabetic retinopathy; ECM, extracellular matrix; ER, endoplasmic reticulum; IL, interleukin; IR, insulin
resistance; FFA, free fatty acid; NET, neutrophil extracellular trap; NF-κB, nuclear factor kappa B; ROS, reactive oxygen
species; SA-β-gal, senescence-associated beta-galactosidase; SAHF, senescence-associated heterochromatin foci;
SASP, senescence-associated secretory phenotype; SnC, senescent cell; T2D, type II diabetes; UPR, unfolded protein
response; VSMC, Vascular smooth muscle cell; WAT, white adipose tissue
Received: 15 April 2021; Editorial Decision: 28 June 2021; First Published Online: 7 August 2021; Corrected and Typeset:
25 August 2021.

Abstract
Cellular senescence is a cell fate that occurs in response to numerous types of stress and
can promote tissue repair or drive inflammation and disruption of tissue homeostasis
depending on the context. Aging and obesity lead to an increase in the senescent cell
burden in multiple organs. Senescent cells release a myriad of senescence-associated
secretory phenotype factors that directly mediate pancreatic β-cell dysfunction, adipose
tissue dysfunction, and insulin resistance in peripheral tissues, which promote the onset
of type II diabetes mellitus. In addition, hyperglycemia and metabolic changes seen in
diabetes promote cellular senescence. Diabetes-induced cellular senescence contributes
to various diabetic complications. Thus, type II diabetes is both a cause and consequence
of cellular senescence. This review summarizes recent studies on the link between aging,
obesity, and diabetes, focusing on the role of cellular senescence in disease processes.
Key Words: Diabetes, cellular senescence, aging, obesity, senotherapeutics, inflammation

Introduction and obesity are the predominant T2D risk factors and are
Type II diabetes (T2D) is a growing public health problem. also associated with an increased burden of senescent cells
Older adults make up the largest segment of people with (SnCs). While cellular senescence is reported to contribute
T2D, and as the number of individuals over 65 is growing to the pathogenesis of diabetes, the diabetic microenviron-
rapidly, so is the incidence of T2D (1). One hundred and ment also seems to increase the burden of SnCs (3).
eleven million cases of T2D were reported in the year 2019 Cellular senescence is a state of indefinite cessation
among people over 65 years of age. In this age group, it of the cell cycle. Senescence can be triggered by mitotic
is estimated that 1 out of 5 people have T2D (2). Aging stress, DNA damage, telomere erosion, mitochondrial

ISSN Online 1945-7170


© The Author(s) 2021. Published by Oxford University Press on behalf of the Endocrine Society. All rights reserved. For
permissions, please e-mail: journals.permissions@oup.com
https://academic.oup.com/endo   1
2 Endocrinology, 2021, Vol. 162, No. 10

dysfunction, oxidative stress, inflammation, mechanical in SnCs plays a positive role in tissue repair, likely through
stress, and oncogenic activation (Fig. 1). Some, but not the secreting the factors that remodel tissue locally and ac-
all, SnCs develop a senescence-associated secretory pheno- tivate the immune system to limit that remodeling (10). In
type (SASP), comprising a variety of factors that include young healthy animals, senescence is induced transiently
proinflammatory interleukins, chemokines, growth factors, upon tissue injury and is essential for the restoration of
proteases, receptors, enzymes that modify the extracellular tissue homeostasis (10-12).
matrix (ECM), stem cell/progenitor toxins, reactive me-
tabolites, bioactive lipids, microRNAs, and extracellular
vesicles. Evidence suggests that the role of SASP is to induce Hallmarks of Cellular Senescence
the immune system to clear the damaged SnCs, since they Cell cycle arrest, mediated by upregulation of the cell
themselves are often resistant to apoptosis. However, with cycle inhibitors p16INK4a (p16), p15INK4b (p15), p19ARF,
and p21CIP (p21), is a universal characteristic of SnCs (Fig.

Downloaded from https://academic.oup.com/endo/article/162/10/bqab136/6345039 by guest on 24 March 2024


age, immune system function wanes contributing to an in-
creased burden of SnCs in older organisms. Senescence is 1). However, it is not a specific marker of senescence, be-
a conserved tumor suppressor mechanism that prevents cause multiple cellular mechanisms can drive cell cycle
replication of a damaged genome and mutagenesis (4). No arrest. Upregulation of cell cycle inhibitors is also not a
known physiologic stimuli can cause SnCs to re-enter the pan-senescence marker since various cell cycle regulators
cell cycle (5, 6). Thus, the cell cycle arrest in SnCs is con- are engaged in different cell types and also in response to
sidered irreversible. Expression of p16INK4a, a cell cycle in- difference inducers of senescence. SnCs do not respond to
hibitor and tumor suppressor, increases exponentially with mitogenic signaling, which distinguishes them from quies-
chronologic age, suggesting that SnCs accumulate more cent cells (13).
rapidly in old age (7), likely due to both an increased in- Most, if not all, SnCs show signs of DNA damage re-
duction and decreased immune clearance (8). Chronic sen- gardless of the mechanism inducing senescence, including
escence is established to play a causative role in aging and increased γ-H2AX foci, 53BP1 foci, telomere-associated
many age-related diseases (9). However, an acute increase foci (TAF), and de-condensation of pericentromeric satellite

Figure 1. Characteristics of senescent cells. SnCs have upregulation of the cell cycle inhibitors p16INK4a and p21CIP1. In addition, SnCs show telomere
shortening or damage (TAF), and epigenetic changes (SAHF). SnCs have decreased levels of nuclear Lamin B1 and increased lysosomal SA-β-gal
activity. Many SnCs have a SASP comprised of chemokines, inflammatory factors and interleukins, growth factors and regulators, extracellular ma-
trix components, soluble receptors, proteases and regulators, reactive metabolites, bioactive lipids, microRNAs, and extracellular vesicles. Early in
senescence, SnCs secrete HMBG1, a key DAMP (an endogenous molecule that activates the innate immune system), which amplifies the SASP. There
is no senescent-specific marker and not all SnCs express the same markers, especially in terms of the SASP. ROS, reactive oxygen species; SA-β-gal,
senescence-associated β-galactosidase; SAHF, senescence-associated heterochromatin foci; SASP, senescence-associated secretory phenotype; TAF,
telomere-associated foci. Figure created with BioRender.com.
Endocrinology, 2021, Vol. 162, No. 10 3

heterochromatin (senescence-associated heterochromatin enriched in epigenetic marks associated with DNA damage
foci) (14) (Fig. 1). DNA damage is a direct inducer of sen- (27). The most consistent change in the plasma membrane
escence through DNA damage response (DDR) signaling composition in SnCs is the upregulation of caveolin-1, an
mechanism (15). For example, double-strand breaks are essential component of cholesterol-enriched microdomains
potent activators of the DDR initiate by identification of called caveolae. Caveolin-1 contributes to the morphology
a broken DNA end by the MRE11/RAD50/NBS1 complex of SnCs (28).
and recruitment of ataxia-telangiectasia mutated (ATM) SnCs have upregulation of many lysosomal proteins and
kinase to that site. ATM functions as a signal transducer, increased content, detected as increased SA-β-gal activity
activating hundreds of effector proteins through its kinase (29) (Fig. 1). Increased lysosomal content could result from
activity to promote DNA repair. This includes phosphor- the metabolic demands of SnCs and enhanced lysosomal
ylation of the histone H2AX, leading to γ-H2AX foci at biogenesis. However, residual bodies, namely lipofuscins

Downloaded from https://academic.oup.com/endo/article/162/10/bqab136/6345039 by guest on 24 March 2024


sites of DNA repair (16). ATM also phosphorylates and in SnCs, support the lack of lysosomal removal (30). SnCs
activates p53, which leads to the transcriptional activa- also have increased mitochondria content. The membrane
tion of many genes that can mediate senescence if the DDR potential of SnC mitochondria is often decreased, leading
signaling persists (17). γ-H2AX nuclear foci or phosphor- to the release of mitochondrial enzymes and increased re-
ylated p53 are commonly used as markers of senescence, active oxygen species (ROS) production (31) (Fig. 1). The
although, again, they are not specific to SnCs. major source of the extra-mitochondrial content is the ac-
As described above, SnCs secrete cytokines, chemokines, cumulation of dysfunctional mitochondria due to reduced
and proteases, which are collectively referred to as the SASPs mitophagy (32). This is, at least in part, a consequence of
(Fig. 1). SASP factors also are not specific to SnCs and can reduced mitochondrial fission and increased fusion (33).
play both positive and negative roles in several biological
processes such as wound healing and cancer progression
(10, 18). The upregulation of SASP factors is mediated by Senotherapeutics
the proinflammatory transcription factor nuclear factor Molecules that can specifically kill SnCs (senolytics)
kappa B (NF-κB) (19). A major inducer of NF-κB activa- or dampen their SASP and other senescence markers
tion is the DDR. GATA4 and C/EBPβ are other transcrip- (senomorphics) have been identified (34). Senolytics often
tion factors involved in the regulation of SASP genes (20, target pro-survival pathways upregulated in SnCs, which
21). However, these transcription factors play myriad roles include BCL-2-BCL-XL, p53-p21, and PI3K-AKT, resulting
in cells and are not unique to SnCs. apoptosis of the SnC. The first senolytic drugs identified
Other characteristics of SnCs include increased protein were the natural product fisetin (35), the combination of
content, partly regulated through increased mTOR activity, dasatinib and quercetin (D + Q) (34), HSP90 inhibitors
increased expression of senescent cell antiapoptotic proteins, (36), and Navitoclax (37). Additional natural products,
and increased activity of lysosomal senescence-associated novel compounds, and re-purposed FDA approved drugs
beta-galactosidase (SA-β-gal) (22). SnCs are metabolically with senolytic activity were identified more recently (38,
active with increased AMP:ATP and ADP:ATP ratios (23). 39). Drugs with senomorphic activity include the mTOR
Multiple conditions, such as oxidative stress, mutations, inhibitor rapamycin, used clinically as an immunosuppres-
infections, and lack of chaperones, can cause endoplasmic sant, and inhibitors of IKK-NF-kB and JAK-STAT (40). In
reticulum (ER) stress, leading to the accumulation and ag- addition, metformin, a drug used to treat T2D, prevents
gregation of proteins. ER stress initiates the unfolded pro- NF-κB activation (41), thereby reducing SASP and acting
tein response (UPR), resulting in reduced protein synthesis, as a senomorphic.
enlarged ER, and export of misfolded proteins. SnCs have
an increased UPR, possibly in response to the increased
protein demand of SASP (24, 25). Senescence Drives Aging
A key morphologic feature of SnCs, at least in vitro, is Various markers of senescence, including SA-β-gal, p16,
an enlarged and irregularly shaped cell body (26) (Fig. 1). TAFs, and activation of the DDR, are increased in tissues
This is driven by cytoskeleton rearrangement, primarily of aged mammals, including rodents (42-44), baboons
vimentin filaments (24). Another hallmark of SnCs is the (45, 46), and humans (47-49), suggesting a link between
loss of lamin B1, a structural protein of the nuclear lamina. SnCs and age-associated pathologies. Studies performed
The destabilization of the nuclear integrity caused by re- using BubR1H/H mice, which age rapidly due to an insuf-
duced lamin B1 results in other nuclear changes, such as ficiency of a pivotal mitotic checkpoint protein, revealed
the loss of condensation of constitutive heterochromatin a causal link between SnCs and aging (50). Prematurely
and the appearance of cytoplasmic chromatin fragments aged tissues in this model, including skeletal muscle, eye,
4 Endocrinology, 2021, Vol. 162, No. 10

and adipose tissue (AT), accumulate high-level of p16Ink4a- the toll-like receptor 4 (TLR4)-MyD88 pathway, which
positive SnCs, which contribute to the sarcopenia, cata- leads to the activation of NF-κB (64, 66). Experiments with
racts, and lipodystrophy. The onset of these disease is isolated human islet cells or with mouse cell lines (MIN6,
delayed in BubR1H/H mice when p16Ink4a-expressing cells INS-1) revealed that TLR4 is required for the secretion of
are ablated via drug-induced activation of transgenic IL-1β following exposure to high FFA concentrations (66).
caspase-8, expression of which is driven by the p16 pro- Knock-down of TLR4 or Myd88 expression, or blocking
motor (referred to as the INK-ATTAC construct) (51). the signaling pathway, attenuates production of IL-1β, and
Elimination of SnCs in aged wild-type INK-ATTAC mice insulin production is restored.
results in an extension in the median but not maximal Numerous studies have documented the accumulation of
lifespan, suggesting that SnCs contribute to mortality in macrophages (MΦ) within islets of T2D patients and mouse
mice. The Ercc1−/Δ mouse model of a human progeroid models of obesity and diabetes (66-69). Infiltration and po-

Downloaded from https://academic.oup.com/endo/article/162/10/bqab136/6345039 by guest on 24 March 2024


syndrome (52) has reduced expression of the ERCC1-XPF larization of MΦ from an anti-inflammatory (M2) profile
endonuclease, which is essential for multiple DNA repair towards a proinflammatory (M1) profile (CD11b+Ly-6C-
mechanisms. Ercc1−/Δ mice spontaneously develop senes- MΦ) are characteristic of islets in T2D (64). Several mouse
cence in multiple organs, including AT, liver, and pancreas, models of obesity, as well as mice placed on a high-fat diet
and interestingly a similar distribution of SnCs is seen (HFD), display a similar increase in islet MΦ. The secretion
in aged wild-type mice (53-55). Removal of SnCs using of chemokines like CCL2 and CXCL1 help to recruit cir-
senolytic drugs blunts multiple aging features in Ercc1−/Δ culating monocytes that convert into CD11b+Ly-6C+ MΦ.
mice and aged wild-type mice, including renal dysfunction, T cells are not increased in islets of mouse models of T2D,
liver damage, pancreatic damage, and neurodegeneration, however, increased B cells are detected (70). In fact, a small
while improving strength and endurance, and activity (9, percentage of T2D patients test positive for autoantibodies
35, 56). Collectively, these studies demonstrate that SnCs to islet proteins (71, 72) (insulin, insulinoma-associated
drive age-related morbidity and mortality (57). protein-2, glutamic acid decarboxylase), which occurs most
frequently in older patients who have had diabetes for a
longer duration. Additionally, T2D patients can also have
Cellular Senescence in Pathogenesis of auto-reactive T cells, which in some elderly patients can
Diabetes occur in the absence of autoantibodies.
The role of SnCs in diabetes is complex. SnCs play a part Glucose tolerance worsens with age in T2D, reflecting
in T2D pathophysiology through a direct effect on pancre- a progressive reduction in the responsiveness of β-cells
atic β-cell function, participation in AT dysfunction, and to glucose stimulation and reduced sensitivity of periph-
SASP-mediated tissue damage (58, 59). Metabolic changes eral tissues to insulin. Prior to the onset of hyperglycemia,
observed in diabetes, such as high circulating glucose and β-cells can compensate for increased metabolic demands
altered lipid metabolism, can also stimulate SnC formation with increased insulin secretion. However, this compen-
(60). Thus, SnCs could be part of a pathogenic loop in dia- sation becomes limited by the age-related decline in β-cell
betes, as both a cause and effect of metabolic changes and proliferation seen in both rodents (73) and humans (74).
tissue damage (58). This lack of proliferation in response to increased insulin
demand may arise partially from the accumulation of sen-
escent β-cells (75).
Pancreatic β-Cells Evidence for senescent β-cells includes diminished
Defective insulin secretion in T2D is caused by β-cell dys- β-cell proliferation and increased expression of senescence
function and reduced β-cell mass. β-cell dysfunction is markers are found in diet-induced T2D mice (76) and the
caused by chronic hyperglycemia and/or hyperlipidemia, telomere length in β-cells from patients with T2D is signifi-
referred to as glucotoxicity and lipotoxicity, respectively cantly shorter than in the control subjects (77). High-glucose
(61). Chronically high levels of free fatty acids (FFAs) and drives oxidative stress in β-cells, which may contribute to
glucose are toxic to β-cells, which can manifest in the form telomerase dysfunction those cells (78). Haploinsufficiency
of ER stress, the production of ROS, and/or mitochondrial of Tert in mice shortens telomeres, including in β-cells.
dysfunction (62, 63). To further complicate matters, β-cells This leads to fasting hyperglycemia, increased expression
respond to FFA by producing and secreting inflammatory of p16Ink4a in pancreatic islets, and impaired mitochon-
cytokines (interleukin [IL]-1β, IL-6, IL-8) and chemokines drial membrane hyperpolarization and Ca2+ mobilization
(CCL2, CXCL1) (64), of which IL-1β is a major player β-cells, which are crucial for insulin exocytosis (79).
and its secretion affects the activation of resident immune The number of SA-β-gal+ cells is increased in islets iso-
cells (65). β-cells mediate their response to FFA through lated from the pancreas of healthy older people compared
Endocrinology, 2021, Vol. 162, No. 10 5

to young, and this is increased further in islets from T2D with obesity is causatively associated with T2D and cardio-
patients compared to nondiabetic, suggesting that β-cell vascular diseases. vWAT is more prone to senescence com-
senescence might contribute to the pathogenesis of T2D pared to sWAT (88). Senescent AT impairs lipid handling,
(80). In mice, SA-β-gal+ β-cells have decreased expres- while promoting IR, defective adaptive thermogenesis,
sion of β-cell identity genes (Ins1, Pdx1, Mafa, Neurod1) aberrant adipocytokine production, and interferes with
with a concomitant increase in expression of markers normal physiological functions of other metabolic organs
of senescence (p16 and p21), and SASP (Ccl2, Il1α, Il6, via SASP.
Tnfα). Senescent β-cells have increased basal insulin se- BubR1H/H mice display dramatic loss of sWAT and
cretion (81) and are transcriptionally rewired, displaying AT senescence (89). Clearance of these p16Ink4a+ cells in
decreased expression of genes required for cellular de- BubR1H/H; INK-ATTAC mice reverses AT senescence,
polarization, incretin signaling, and production of insulin thereby, preventing aging-related lipodystrophy (9).

Downloaded from https://academic.oup.com/endo/article/162/10/bqab136/6345039 by guest on 24 March 2024


granules (82), all necessary for insulin secretion. AT from obese individuals exhibits increased oxidative
Senescent β-cells have also occurred in the pancreas stress and telomere shortening, which induce senescence.
of patients with type 1 autoimmune diabetes (T1D) com- Intracellular ROS activates p38 mitogen-activated protein
pared with nondiabetic people and individuals who are kinases, which in turn induces p53-p21–dependent senes-
autoantibody positive prior to onset of T1D (83). This cence (87). IR is one of the consequences of AT senescence
suggests a correlation between disease progression and the when p53 is activated. Transgenic mice with overexpression
level of senescent β-cells. In the nonobese diabetic mouse, of p53 have increased senescence and IR, whereas p53 in-
an animal model of T1D, SnCs and SASP increase with activation results in the opposite effects in diabetic mice
time in pancreatic islets (83). Senolytics reduce senescence (59, 90).
markers and the incidence of diabetes in these mice (83). Metabolic stress induced by high calorie intake in-
In mice placed on an HFD, the onset of insulin resistance duces an inflammatory microenvironment, in particular in
(IR) is associated with senescent β-islet cells and senolytics AT. This affects not only resident immune cells, but also
can restore glucose metabolism (84). p16INK4a expression infiltrating immune cells, which heighten inflammation
increases in human and mouse β-cells with age, and this and contribute to IR through secretion of inflammatory
contributes to their reduction in regenerative capacity cytokines (82, 91-93). Excessive nutrient intake activates
(79, 85, 86). Clearance of p16Ink4a+ cells in mice, using adipocytes to internalize and accumulate FFA through
the INK-ATTAC construct, improves glucose metabolism a CD36-driven mechanism regulated by PPARγ (94-
and insulin secretion, while decreasing the expression of 96). Adipocytes respond to the accumulation of lipids by
proinflammatory SASP factors in islets (80). The senolytic enlarging and proliferating and initiating an inflammatory
ABT263 (Navitoclax), reverses hyperglycemia and restores response through an HIF1α-NF-kB–driven mechanism (97-
the normal β-cell gene expression profile mice with induced 99). Similar to adipocytes, adipose resident MΦ also accu-
insulin resistance caused by the insulin receptor antagonist mulate FFA through CD36. These MΦ further contribute to
S961 (80). The BCL-2 pathway, 1 of several anti-apoptotic the inflammatory response by secreting TNF-α, IL-6, Il-1β,
pathways, is upregulated in SA-β‐gal+ β‐cells, consistent MCP-1, CX3CL1, and IL-8 (94, 95, 100). vWAT of wild-
with ABT263 killing SA-β‐gal+ islet cells in vitro (80). type and db/db mice on a HFD exhibit elevated markers of
senescence (SA-β-gal, p16Ink4a) (101), a diminished capacity
to clear glucose upon challenge, and IR. Senolytics signifi-
Adipose Tissue Senescence and Inflammation cantly improve glucose tolerance and insulin sensitivity
AT is the largest and most active endocrine organ for (101). Exercise has a similar effect as senolytics in these
regulating energy balance, glucose and lipid homeostasis, diabetic mice (102).
and immunity in response to the fluctuation in nutritional There are numerous interactions between the immune
status, environment, lifestyle, and aging (87). White adipose system and AT during obesity that contribute to IR (Table
tissue (WAT) stores surplus energy in the form of triglycer- 1). This inflammatory response, and the subsequent hyper-
ides that are hydrolyzed into FFA and glycerol for energy glycemia, enhances the formation of SnCs, which can ex-
supply during nutrient deprivation. Brown AT (BAT) or acerbate IR (83, 84). The number of senescent adipocytes
beige adipocytes convert chemical energy, mainly FFA, into increases the number of MΦ positive for the cyclic ADP
heat via the mitochondrial uncoupling reaction to maintain ribose hydrolase CD38 (103). CD38 expression increases
core body temperature and energy balance. The amount along with inflammatory and senescence markers in the
and function of AT are closely associated with age, calorie AT of old mice, in addition to CD68, a marker of active
intake, physical activity, and health status. An increase in MΦ (103). Increased CD38 drives depletion of the critical
visceral WAT (vWAT), but not subcutaneous WAT (sWAT) metabolite NAD+ (104). The increase in IR is paralleled
6 Endocrinology, 2021, Vol. 162, No. 10

Table 1. A short list of the interactions occurring between immune cells and inflamed adipose tissue

Immune cell subsets Impact on adipose tissue

T cells Peripheral CD8+ T cells infiltrate inflamed AT (92, 93, 100, 136)
Infiltrating T cells adopt a Th1 (interferon secreting) or a Th17 (IL-17-secreting) profile (93, 137-139)
Decreased antigen specificities recognized by infiltrating CD8+ T cells (100)
Resident CD4+ regulatory T cells maintain AT homeostasis through PPARγ. However, during obesity, the frequency
of resident CD4+ regulatory T cells decreases (136, 140-143)
Macrophage Infiltration into AT mediated by MCP-1 chemokine secreted by resident T cells (97, 136)
Excessive nutrients induce activation of MΦ driven through HIF1α, NF-κB, and IL-1β (94, 95, 136, 144)
DC Infiltration into AT mediated by CX3CL1 chemokine secreted by resident T cells (97, 145)
In lean mice, DCs exhibit a CD11c+CD103+ phenotype that induces CD4+Foxp3+ Tregs to maintain AT homeostasis.
In obese or db/db knockout mice, DC exhibit a CD11c+F4/80lo phenotype and induce Th17 cells (145, 146)

Downloaded from https://academic.oup.com/endo/article/162/10/bqab136/6345039 by guest on 24 March 2024


Other innate cells In lean conditions, innate lymphoid cells type-2 induce anti-inflammatory responses and attenuate insulin resistance,
but this is lost during obesity (147-149)

Abbreviations: AT, adipose tissue; DC, dendritic cell; IL, interleukin; NF-κB, nuclear factor kappa B.

by an increase in senescent T cells both in humans and Cellular Senescence in Diabetic Complications
mice (105). Older T2D patients have elevated levels of
People with diabetes are more likely to develop age-related
CD8+CD57+CD28– T cells, which exhibit markers of sen-
morbidities, such as frailty, Alzheimer’s disease, mild cog-
escence. Levels of senescent CD8+ T cells are increased in
nitive impairment, cardiovascular disease, osteoporosis,
pre-diabetics compared to nondiabetic controls and may be
visual impairment, and bladder and renal dysfunction,
predictive of the development of hyperglycemia (105, 106).
indicating that T2D itself might represent a proaging state
Senescent T cells in mice, defined as CD8+CD44+PD-1+,
(3). Advanced glycation end products (AGEs) are pro-
are elevated in old wild-type mice placed on a HFD for
teins or lipids that become glycated when combined with
2 months, which leads to IR, inflamed AT, and increased
sugars. AGEs contribute to multiple microvascular and
SnCs in AT and liver (105). Notably, driving senescence in
macrovascular complications by forming of cross-links be-
the immune system alone, is sufficient to drive increased
tween molecules in the basement membrane of the ECM
expression of senescence markers in pancreatic tissue and
and by engaging the receptor for advanced glycation end
organ damage, as determined by serum amylase levels (107).
products (111). In diabetes, the increased AGE formation
induces endothelial cell senescence (112). Ex vivo, endo-
thelial cells from healthy aged and T2D individuals show
Peripheral Tissues chronic NF-κB activation. Blockade of NF-κB-mediated
Fat is redistributed outside fat depots with aging, accumu- inflammatory responses in endothelial cells prolongs the
lating in muscle, liver, bone marrow, and other ectopic sites. lifespan of mice by preventing IR (113). Endothelial cell
Insulin-resistant hypertrophic adipocytes have augmented senescence impairs systemic metabolic health (114) (Fig.
lipolytic activity and diminished ability to take up FFA, re- 2). The presence of elevated oxidized low-density lipopro-
sulting in a redirection of lipids toward nonadipose tissues teins, commonly observed in diabetics, reduces the number
(ectopic fat deposition) (108). If fatty acid β-oxidation in the of circulating endothelial progenitor cells and drives their
mitochondria cannot keep up with the increased supply of senescence (115).
FFA to nonadipose tissues accumulation of lipid intermedi-
ates like diacylglycerol and ceramides occurs. These lipid
intermediates lead to activation of serine/threonine kinases Diabetic Retinopathy
that phosphorylate insulin receptor substrates on serine In diabetic retinopathy (DR), the microvessels in the
residues. Phosphorylated insulin receptor substrates do not retina become damaged, degenerate, and regrow in an
function properly, leading to impaired insulin signaling and aberrant manner. Diseased vessels show upregulation
disruption of normal metabolic processes. Accumulation of of senescence marker expression (116). In retinas of pa-
lipid intermediates promotes ER stress, mitochondrial dys- tients with DR and in a mouse model of proliferative
function, and apoptosis (109). Dysregulation of liver and DR, a high level of senescent p16INK4a-expressing cells
skeletal muscle metabolism can strongly impact whole- accumulates. Senolysis by means of genetic approaches
body glucose homeostasis and insulin sensitivity (110). that clear p16Ink4a expressing cells or small molecule
Endocrinology, 2021, Vol. 162, No. 10 7

Downloaded from https://academic.oup.com/endo/article/162/10/bqab136/6345039 by guest on 24 March 2024


Figure 2. Diabetes and cellular senescence. Aging and obesity induce inflammation and cellular senescence in pancreatic β-cells and adipocytes.
Senescence of β-cells results in β-cell dysfunction and decreased insulin exocytosis. Senescence of adipocytes leads to ectopic lipid accumulation
and insulin resistance. Together these contribute to progression of T2D. Hyperglycemia-induced endothelial senescence plays a major role in devel-
opment of various diabetic complications such as diabetic retinopathy, diabetic nephropathy, and cardiovascular complications. Figure created with
BioRender.com.

inhibitors of the antiapoptotic protein BCL-xL, sup- the production of neutrophil extracellular traps (NETs).
press pathological angiogenesis (116). In DR, endothelial NETs ultimately clear senescent endothelial cells and re-
cells are the primary targets of glucose-induced damage. model unhealthy blood vessels. Genetic or pharmacological
Diabetes-induced retinal endothelial cell senescence in- inhibition of NETosis prevents the clearance of senescent
volves sequential events initiated by NADPH oxidase-2 endothelial cells and promotes DR (124).
(NOX2) activation, increased expression of its target
arginase-1 (Arg1), and ROS production (117). Increased
Arg1 expression promotes senescence through a mech- Cardiovascular Complications
anism involving increases in the expression of p16Ink4a, Vascular aging contributes to the high morbidity and
p21Cip1, and p53 (118). Arg1 gene deletion or pharma- mortality of cardiovascular diseases observed in patients
cology inhibition protects against endothelial cell senes- with diabetes. Vascular smooth muscle cells (VSMCs) are
cence in diabetic mice (118). vital components of the vascular walls and senescence of
Glucose-induced downregulation of SIRT1 in endothe- these cells plays a key role in vascular aging. For instance,
lial cells promotes senescence and increases the production VSMC senescence can promote vascular calcification and
of several ECM proteins and vasoactive factors in diabetes atherosclerosis (125). Vascular calcification is a key com-
(119). MicroRNAs (miR-34a, miR-1, miR-19b, miR-320a) ponent of vasculopathy commonly seen in patients with
are translational suppressors of SIRT1 and also implicated T2D. SnCs in proximity to blood vessel walls (eg, in peri-
in endothelial cell senescence (120, 121). Oxidative stress vascular AT) may contribute to the cellular dysfunction
suppression of SIRT6 activity also drives endothelial cell that underlies the development and progression of vas-
senescence and likely contributes to the pathogenesis of DR cular diseases such as aneurysm and atherosclerosis (126).
(122). Endothelial cell senescence is reduced by Donepezil, In VSMCs, adiponectin can reduce high-glucose-induced
which increases SIRT1 activity (123). Senescent endothelial senescence and may be a potential therapeutic agent for
cells have a secretome that attracts neutrophils and triggers cardiovascular complications in diabetic patients (127).
8 Endocrinology, 2021, Vol. 162, No. 10

Diabetic Nephropathy side effects have been observed to date (56). Therefore, it
Diabetic nephropathy is the leading cause of chronic kidney is possible that senolytics, once optimized, will be effective
disease and renal failure in developed countries. In the past two in delaying the onset and reducing the severity of diabetic
decades, the morbidity and mortality of diabetic nephropathy complications, particularly in obese individuals.
have risen rapidly globally (128). The mechanisms driving dia-
betic nephropathy involve a multifactorial interaction of meta-
bolic and hemodynamic factors, including high blood glucose,
Acknowledgements
AGE, the renin–angiotensin system, and protein kinase-C–in- Financial Support: This work was supported by the NIH/NIA
grants P01 R01 AG063543, R01 AG043376, U19 AG056278,
duced generation of ROS, which mediates the activation of
and P01 AG062413.
the transcription factor NF-κB (129). Diabetic nephropathy Conflicts of Interest: Pending patents on senolytic drugs and their
is strongly associated with accelerated senescence of renal uses are held the University of Minnesota (L.J.N., P.D.R.). L.J.N. and

Downloaded from https://academic.oup.com/endo/article/162/10/bqab136/6345039 by guest on 24 March 2024


tubular cells, podocytes, mesangial cells, and endothelial cells. P.D.R. are co-founders of NRTK Biosciences, a startup focused on
Hyperglycemia can directly induce senescence in mesangial the development of novel senolytics. L.J.N. has consulted for Merck
and Ono Pharma on senescent cells as a therapeutic target. L.J.N. and
(130) and tubular cells (131). Interestingly, high glucose can
P.D.R. are on the Scientific Advisory Board for Innate Biologics,
also induce MΦ to secrete inflammatory factors, creating a
developing bacterial effector proteins, and P.D.R. is co-founder and
state of low-grade inflammation (129). member of the Scientific Advisory Board for Genascence Corporation,
In both T1D and T2D, telomere attrition is associated a gene therapy company focused on osteoarthritis.
with renal cell senescence, proteinuria, and progression of
diabetic nephropathy (132, 133). In streptozotocin-induced
T1D mice, renal expression of p21 is increased dramatic- Additional Information
ally along with an increase in SA-β-gal staining in tubular Correspondence: Laura J. Niedernhofer, MD, PhD, Institute on
epithelial cells. However, other cyclin-dependent kinase the Biology of Aging and Metabolism, Department of Biochemistry,
inhibitors such as p16Ink4a and p27Kip1 were not altered. Molecular Biology and Biophysics, University of Minnesota Medical
School, 6-155 Jackson Hall, 321 Church Street, SE, Minneapolis,
In the T1D kidney, tubular senescence was induced by
MN 55455, USA. Email: lniedern@umn.edu.
hyperglycemia via a sodium-glucose transport protein 2 Data Availability: Data sharing is not applicable to this art-
(SGLT2) (134). Glomerular mesangial cells, which provide icle as no datasets were generated or analyzed during the current
structural support for the glomerular capillary loops and study.
regulate the ultrafiltration surface for glomerular filtration,
are directly affected by hyperglycemia. A high extracellular
glucose concentration promotes ECM accumulation, cell
References
cycle arrest, cellular hypertrophy, and induces senescence
1. Kalyani RR, Golden SH, Cefalu WT. Diabetes and aging:
in cultured glomerular mesangial cells (135).
unique considerations and goals of care. Diabetes Care.
2017;40(4):440-443.
2. Khan RMM, Chua ZJY, Tan JC, Yang Y, Liao Z, Zhao Y. From
Conclusion
pre-diabetes to diabetes: diagnosis, treatments and translational
The accumulation of SnCs during aging and obesity is research. Medicina (Kaunas) 2019;55(9):546.
involved in the pathogenesis of T1D and T2D. High glu- 3. Palmer AK, Gustafson B, Kirkland JL, Smith U. Cellular senes-
cose induces senescence of a variety of cell types, which cence: at the nexus between ageing and diabetes. Diabetologia.
2019;62(10):1835-1841.
contributes to the development of diabetic complications,
4. van Deursen JM. The role of senescent cells in ageing. Nature.
including nephropathy, retinopathy, vasculopathy, and
2014;509(7501):439-446.
cardiovascular disease. Thus, SnCs play a major role in 5. Regulski MJ. Cellular Senescence: what, why, and how. Wounds.
initiating and exacerbating diabetes. Given that oral ad- 2017;29(6):168-174.
ministration of senolytic combination dasatinib and quer- 6. Hernandez-Segura A, Nehme J, Demaria M. Hallmarks of
cetin (D+Q) improves insulin sensitivity and reduces the Cellular Senescence. Trends Cell Biol. 2018;28(6):436-453.
AT inflammation in obese mice (101), 2 phase 2, random- 7. Liu Y, Sanoff HK, Cho H, et al. Expression of p16(INK4a) in
ized clinical trials of D+Q and the senolytic fisetin were peripheral blood T-cells is a biomarker of human aging. Aging
Cell. 2009;8(4):439-448.
initiated for subjects with diabetic kidney disease at Mayo
8. He S, Sharpless NE. Senescence in Health and Disease. Cell.
Clinic (NCT03325322 and NCT02848131). For the D+Q
2017;169(6):1000-1011.
study, senolytic treatment has already been demonstrated 9. Baker DJ, Wijshake T, Tchkonia T, et al. Clearance of p16Ink4a-
to reduce the senescent cell burden in AT biopsied from positive senescent cells delays ageing-associated disorders.
obese diabetic subjects (56). Importantly, no serious drug Nature. 2011;479(7372):232-236.
Endocrinology, 2021, Vol. 162, No. 10 9

10. Demaria M, Ohtani N, Youssef SA, et al. An essential role for 28. Zou H, Stoppani E, Volonte D, Galbiati F. Caveolin-1, cel-
senescent cells in optimal wound healing through secretion of lular senescence and age-related diseases. Mech Ageing Dev.
PDGF-AA. Dev Cell. 2014;31(6):722-733. 2011;132(11-12):533-542.
11. Storer M, Mas A, Robert-Moreno A, et al. Senescence is a devel- 29. Lee BY, Han JA, Im JS, et al. Senescence-associated beta-
opmental mechanism that contributes to embryonic growth and galactosidase is lysosomal beta-galactosidase. Aging Cell.
patterning. Cell. 2013;155(5):1119-1130. 2006;5(2):187-195.
12. Collado M, Serrano M. Senescence in tumours: evidence from 30. Georgakopoulou EA, Tsimaratou K, Evangelou K, et al.
mice and humans. Nat Rev Cancer. 2010;10(1):51-57. Specific lipofuscin staining as a novel biomarker to detect rep-
13. Dimri GP, Campisi J. Molecular and cell biology of replicative licative and stress-induced senescence. A method applicable
senescence. Cold Spring Harb Symp Quant Biol. 1994;59:67-73. in cryo-preserved and archival tissues. Aging (Albany NY).
14. Bernadotte A, Mikhelson VM, Spivak IM. Markers of cellular 2013;5(1):37-50.
senescence. Telomere shortening as a marker of cellular senes- 31. Passos JF, Saretzki G, Ahmed S, et al. Mitochondrial dysfunction
cence. Aging (Albany NY). 2016;8(1):3-11. accounts for the stochastic heterogeneity in telomere-dependent

Downloaded from https://academic.oup.com/endo/article/162/10/bqab136/6345039 by guest on 24 March 2024


15. White RR, Vijg J. Do DNA Double-Strand Breaks Drive Aging? senescence. PLoS Biol. 2007;5(5):e110.
Mol Cell. 2016;63(5):729-738. 32. Tai H, Wang Z, Gong H, et al. Autophagy impairment with
16. Celeste A, Petersen S, Romanienko PJ, et al. Genomic instability in lysosomal and mitochondrial dysfunction is an important char-
mice lacking histone H2AX. Science. 2002;296(5569):922-927. acteristic of oxidative stress-induced senescence. Autophagy.
17. Turenne GA, Paul P, Laflair L, Price BD. Activation of p53 2017;13(1):99-113.
transcriptional activity requires ATM’s kinase domain and 33. Dalle Pezze P, Nelson G, Otten EG, et al. Dynamic modelling of
multiple N-terminal serine residues of p53. Oncogene. pathways to cellular senescence reveals strategies for targeted
2001;20(37):5100-5110. interventions. PLoS Comput Biol. 2014;10(8):e1003728.
18. Coppé JP, Desprez PY, Krtolica A, Campisi J. The senescence- 34. Zhu Y, Tchkonia T, Pirtskhalava T, et al. The Achilles’ heel of
associated secretory phenotype: the dark side of tumor suppres- senescent cells: from transcriptome to senolytic drugs. Aging
sion. Annu Rev Pathol. 2010;5:99-118. Cell. 2015;14(4):644-658.
19. Ohanna M, Giuliano S, Bonet C, et al. Senescent cells develop 35. Yousefzadeh MJ, Zhu Y, McGowan SJ, et al. Fisetin is a
a PARP-1 and nuclear factor-{kappa}B-associated secretome senotherapeutic that extends health and lifespan. Ebiomedicine.
(PNAS). Genes Dev. 2011;25(12):1245-1261. 2018;36:18-28.
20. Kuilman T, Michaloglou C, Vredeveld LC, et al. Oncogene- 36. Schafer MJ, White TA, Iijima K, et al. Cellular senescence medi-
induced senescence relayed by an interleukin-dependent inflam- ates fibrotic pulmonary disease. Nat Commun. 2017;8:14532.
matory network. Cell. 2008;133(6):1019-1031. 37. Chang J, Wang Y, Shao L, et al. Clearance of senescent cells by
21. Kang C, Xu Q, Martin TD, et al. The DNA damage response ABT263 rejuvenates aged hematopoietic stem cells in mice. Nat
induces inflammation and senescence by inhibiting autophagy Med. 2016;22(1):78-83.
of GATA4. Science 2015;349(6255):aaa5612. 38. Robbins PD, Jurk D, Khosla S, et al. Senolytic drugs: redu-
22. Childs BG, Baker DJ, Kirkland JL, Campisi J, van Deursen JM. cing senescent cell viability to extend health span. Annu Rev
Senescence and apoptosis: dueling or complementary cell fates? Pharmacol Toxicol. 2021;61:779-803.
EMBO Rep. 2014;15(11):1139-1153. 39. Borghesan M, Hoogaars WMH, Varela-Eirin M, Talma N,
23. James EL, Michalek RD, Pitiyage GN, et al. Senescent human Demaria M. A senescence-centric view of aging: implications for
fibroblasts show increased glycolysis and redox homeostasis longevity and disease. Trends Cell Biol. 2020;30(10):777-791.
with extracellular metabolomes that overlap with those of ir- 40. Xu M, Palmer AK, Ding H, et al. Targeting senescent cells en-
reparable DNA damage, aging, and disease. J Proteome Res. hances adipogenesis and metabolic function in old age. Elife.
2015;14(4):1854-1871. 2015;4:e12997.
24. Druelle C, Drullion C, Deslé J, et al. ATF6α regulates morpho- 41. Kanigur Sultuybek G, Soydas T, Yenmis G. NF-κB as the medi-
logical changes associated with senescence in human fibro- ator of metformin’s effect on ageing and ageing-related diseases.
blasts. Oncotarget. 2016;7(42):67699-67715. Clin Exp Pharmacol Physiol. 2019;46(5):413-422.
25. Pluquet O, Pourtier A, Abbadie C. The unfolded protein re- 42. Burd CE, Sorrentino JA, Clark KS, et al. Monitoring tumorigen-
sponse and cellular senescence. A review in the theme: cellular esis and senescence in vivo with a p16(INK4a)-luciferase model.
mechanisms of endoplasmic reticulum stress signaling in health Cell. 2013;152(1-2):340-351.
and disease. Am J Physiol Cell Physiol. 2015;308(6):C415 43. Krishnamurthy J, Torrice C, Ramsey MR, et al. Ink4a/Arf expres-
-C425. sion is a biomarker of aging. J Clin Invest. 2004;114(9):1299-1307.
26. Bent EH, Gilbert LA, Hemann MT. A senescence secretory 44. Wang C, Jurk D, Maddick M, Nelson G, Martin-Ruiz C,
switch mediated by PI3K/AKT/mTOR activation controls von Zglinicki T. DNA damage response and cellular senescence
chemoprotective endothelial secretory responses. Genes Dev. in tissues of aging mice. Aging Cell. 2009;8(3):311-323.
2016;30(16):1811-1821. 45. Herbig U, Ferreira M, Condel L, Carey D, Sedivy JM. Cellular
27. Sadaie M, Salama R, Carroll T, et al. Redistribution of the senescence in aging primates. Science. 2006;311(5765):1257.
Lamin B1 genomic binding profile affects rearrangement of het- 46. Jeyapalan JC, Ferreira M, Sedivy JM, Herbig U. Accumulation
erochromatic domains and SAHF formation during senescence. of senescent cells in mitotic tissue of aging primates. Mech
Genes Dev. 2013;27(16):1800-1808. Ageing Dev. 2007;128(1):36-44.
10  Endocrinology, 2021, Vol. 162, No. 10

47. Melk A, Schmidt BM, Takeuchi O, Sawitzki B, Rayner DC, 67. Chan JY, Lee K, Maxwell EL, Liang C, Laybutt DR. Macrophage
Halloran PF. Expression of p16INK4a and other cell cycle regu- alterations in islets of obese mice linked to beta cell disruption
lator and senescence associated genes in aging human kidney. in diabetes. Diabetologia. 2019;62(6):993-999.
Kidney Int. 2004;65(2):510-520. 68. Ehses JA, Perren A, Eppler E, et al. Increased number of
48. Ressler S, Bartkova J, Niederegger H, et al. p16INK4A is a ro- islet-associated macrophages in type 2 diabetes. Diabetes.
bust in vivo biomarker of cellular aging in human skin. Aging 2007;56(9):2356-2370.
Cell. 2006;5(5):379-389. 69. Richardson SJ, Willcox A, Bone AJ, Foulis AK, Morgan NG.
49. Chkhotua AB, Gabusi E, Altimari A, et al. Increased expres- Islet-associated macrophages in type 2 diabetes. Diabetologia.
sion of p16(INK4a) and p27(Kip1) cyclin-dependent kinase 2009;52(8):1686-1688.
inhibitor genes in aging human kidney and chronic allograft 70. Butcher MJ, Hallinger D, Garcia E, et al. Association of
nephropathy. Am J Kidney Dis. 2003;41(6):1303-1313. proinflammatory cytokines and islet resident leucocytes
50. Hanks S, Coleman K, Reid S, et al. Constitutional aneuploidy with islet dysfunction in type 2 diabetes. Diabetologia.
and cancer predisposition caused by biallelic mutations in 2014;57(3):491-501.

Downloaded from https://academic.oup.com/endo/article/162/10/bqab136/6345039 by guest on 24 March 2024


BUB1B. Nat Genet. 2004;36(11):1159-1161. 71. Li R, Huang J, Yu Y, Yang Y. Islet autoantibody patterns in pa-
51. Baker DJ, Perez-Terzic C, Jin F, et al. Opposing roles for tients with type 2 diabetes aged 60 and higher: a cross-sectional
p16Ink4a and p19Arf in senescence and ageing caused by study in a Chinese hospital. Front Endocrinol (Lausanne).
BubR1 insufficiency. Nat Cell Biol. 2008;10(7):825-836. 2018;9:260.
52. Niedernhofer LJ, Garinis GA, Raams A, et al. A new progeroid 72. Brooks-Worrell B, Narla R, Palmer JP. Islet autoimmunity in
syndrome reveals that genotoxic stress suppresses the phenotypic type 2 diabetes patients. Diabetes Obes Metab.
somatotroph axis. Nature. 2006;444(7122):1038-1043. 2013;15(Suppl 3):137-140.
53. Gregg SQ, Robinson AR, Niedernhofer LJ. Physiological con- 73. Teta M, Long SY, Wartschow LM, Rankin MM, Kushner JA.
sequences of defects in ERCC1-XPF DNA repair endonuclease. Very slow turnover of beta-cells in aged adult mice. Diabetes.
DNA Repair (Amst). 2011;10(7):781-791. 2005;54(9):2557-2567.
54. Robinson AR, Yousefzadeh MJ, Rozgaja TA, et al. Spontaneous 74. Gregg BE, Moore PC, Demozay D, et al. Formation of a human
DNA damage to the nuclear genome promotes senescence, β-cell population within pancreatic islets is set early in life. J
redox imbalance and aging. Redox Biol. 2018;17:259-273. Clin Endocrinol Metab. 2012;97(9):3197-3206.
55. Yousefzadeh MJ, Zhao J, Bukata C, et al. Tissue specificity of 75. Li N, Liu F, Yang P, et al. Aging and stress induced β cell sen-
senescent cell accumulation during physiologic and accelerated escence and its implication in diabetes development. Aging
aging of mice. Aging Cell. 2020;19(3):e13094. (Albany NY). 2019;11(21):9947-9959.
56. Xu M, Pirtskhalava T, Farr JN, et al. Senolytics improve 76. Sone H, Kagawa Y. Pancreatic beta cell senescence contributes
physical function and increase lifespan in old age. Nat Med. to the pathogenesis of type 2 diabetes in high-fat diet-induced
2018;24(8):1246-1256. diabetic mice. Diabetologia. 2005;48(1):58-67.
57. Childs BG, Gluscevic M, Baker DJ, et al. Senescent cells: an 77. Tamura Y, Izumiyama-Shimomura N, Kimbara Y, et al.
emerging target for diseases of ageing. Nat Rev Drug Discov. β-cell telomere attrition in diabetes: inverse correlation be-
2017;16(10):718-735. tween HbA1c and telomere length. J Clin Endocrinol Metab.
58. Palmer AK, Tchkonia T, LeBrasseur NK, Chini EN, Xu M, 2014;99(8):2771-2777.
Kirkland JL. Cellular senescence in type 2 diabetes: a thera- 78. Matsui-Hirai H, Hayashi T, Yamamoto S, et al. Dose-dependent
peutic opportunity. Diabetes. 2015;64(7):2289-2298. modulatory effects of insulin on glucose-induced endothelial
59. Tchkonia T, Morbeck DE, Von Zglinicki T, et al. Fat tissue, senescence in vitro and in vivo: a relationship between telomeres
aging, and cellular senescence. Aging Cell. 2010;9(5):667-684. and nitric oxide. J Pharmacol Exp Ther. 2011;337(3):591-599.
60. Regulski M. Understanding diabetic induction of cellular senes- 79. Guo N, Parry EM, Li LS, et al. Short telomeres compromise
cence: a concise review. Wounds. 2018;30(4):96-101. β-cell signaling and survival. PLoS One. 2011;6(3):e17858.
61. Robertson RP, Harmon J, Tran PO, Poitout V. Beta-cell glucose 80. Aguayo-Mazzucato C, Andle J, Lee TB Jr, et al. Acceleration of
toxicity, lipotoxicity, and chronic oxidative stress in type 2 dia- β cell aging determines diabetes and senolysis improves disease
betes. Diabetes. 2004;53(Suppl 1):S119-S124. outcomes. Cell Metab. 2019;30(1):129-142.e4.
62. Fonseca SG, Gromada J, Urano F. Endoplasmic reticulum 81. Aguayo-Mazzucato C, van Haaren M, Mruk M, et al. β Cell
stress and pancreatic β-cell death. Trends Endocrinol Metab. aging markers have heterogeneous distribution and are induced
2011;22(7):266-274. by insulin resistance. Cell Metab. 2017;25(4):898-910.e5.
63. Lytrivi M, Castell AL, Poitout V, Cnop M. Recent insights into 82. Aguayo-Mazzucato C. Functional changes in beta cells during
mechanisms of β-cell lipo- and glucolipotoxicity in type 2 dia- ageing and senescence. Diabetologia. 2020;63(10):2022-2029.
betes. J Mol Biol. 2020;432(5):1514-1534. 83. Thompson PJ, Shah A, Ntranos V, Van Gool F, Atkinson M,
64. Eguchi K, Nagai R. Islet inflammation in type 2 diabetes and Bhushan A. Targeted elimination of senescent beta cells prevents
physiology. J Clin Invest. 2017;127(1):14-23. type 1 diabetes. Cell Metab. 2019;29(5):1045-1060.e10.
65. Maedler K, Sergeev P, Ris F, et al. Glucose-induced beta cell pro- 84. Aguayo-Mazzucato C, Andle J, Lee TB Jr, et al. Acceleration of
duction of IL-1beta contributes to glucotoxicity in human pan- β cell aging determines diabetes and senolysis improves disease
creatic islets. J Clin Invest. 2002;110(6):851-860. outcomes. Cell Metab. 2019;30(1):129-142.e4.
66. Eguchi K, Manabe I, Oishi-Tanaka Y, et al. Saturated fatty acid 85. Krishnamurthy J, Ramsey MR, Ligon KL, et al. p16INK4a in-
and TLR signaling link β cell dysfunction and islet inflamma- duces an age-dependent decline in islet regenerative potential.
tion. Cell Metab. 2012;15(4):518-533. Nature. 2006;443(7110):453-457.
Endocrinology, 2021, Vol. 162, No. 10 11

86. Chen H, Gu X, Su IH, et al. Polycomb protein Ezh2 regulates 104. Chini CCS, Peclat TR, Warner GM, et al. CD38 ecto-enzyme in
pancreatic beta-cell Ink4a/Arf expression and regeneration in immune cells is induced during aging and regulates NAD+ and
diabetes mellitus. Genes Dev. 2009;23(8):975-985. NMN levels. Nat Metab. 2020;2(11):1284-1304.
87. Liu Z, Wu KKL, Jiang X, Xu A, Cheng KKY. The role of adi- 105. Yi HS, Kim SY, Kim JT, et al. T-cell senescence contributes to
pose tissue senescence in obesity- and ageing-related metabolic abnormal glucose homeostasis in humans and mice. Cell Death
disorders. Clin Sci (Lond). 2020;134(2):315-330. Dis. 2019;10(3):249.
88. Yamada T, Kamiya M, Higuchi M, Nakanishi N. Fat depot- 106. Lau EYM, Carroll EC, Callender LA, et al. Type 2 diabetes is
specific differences of macrophage infiltration and cellular associated with the accumulation of senescent T cells. Clin Exp
senescence in obese bovine adipose tissues. J Vet Med Sci. Immunol. 2019;197(2):205-213.
2018;80(10):1495-1503. 107. Yousefzadeh MJ, Flores RR, Zhu Y, et al. An aged immune
89. Baker DJ, Jeganathan KB, Cameron JD, et al. BubR1 insuffi- system drives senescence and ageing of solid organs. Nature.
ciency causes early onset of aging-associated phenotypes and 2021;594(7861):100-105.
infertility in mice. Nat Genet. 2004;36(7):744-749. 108. Blüher M. Adipose tissue dysfunction in obesity. Exp Clin

Downloaded from https://academic.oup.com/endo/article/162/10/bqab136/6345039 by guest on 24 March 2024


90. Minamino T, Orimo M, Shimizu I, et al. A crucial role for adi- Endocrinol Diabetes. 2009;117(6):241-250.
pose tissue p53 in the regulation of insulin resistance. Nat Med. 109. Abdul-Ghani MA, DeFronzo RA. Pathogenesis of in-
2009;15(9):1082-1087. sulin resistance in skeletal muscle. J Biomed Biotechnol.
91. Ehrhardt N, Cui J, Dagdeviren S, et al. Adiposity-independent 2010;2010:476279.
effects of aging on insulin sensitivity and clearance in mice and 110. Petersen KF, Shulman GI. Pathogenesis of skeletal muscle in-
humans. Obesity (Silver Spring). 2019;27(3):434-443. sulin resistance in type 2 diabetes mellitus. Am J Cardiol.
92. Daryabor G, Atashzar MR, Kabelitz D, Meri S, Kalantar K. The 2002;90(5A):11G-18G.
effects of type 2 diabetes mellitus on organ metabolism and the 111. Goldin A, Beckman JA, Schmidt AM, Creager MA. Advanced
immune system. Front Immunol. 2020;11:1582. glycation end products: sparking the development of diabetic
93. Lee BC, Lee J. Cellular and molecular players in adipose tissue vascular injury. Circulation. 2006;114(6):597-605.
inflammation in the development of obesity-induced insulin re- 112. Nowotny K, Jung T, Höhn A, Weber D, Grune T. Advanced
sistance. Biochim Biophys Acta. 2014;1842(3):446-462. glycation end products and oxidative stress in type 2 diabetes
94. Chawla A, Barak Y, Nagy L, Liao D, Tontonoz P, Evans RM. mellitus. Biomolecules. 2015;5(1):194-222.
PPAR-gamma dependent and independent effects on 113. Hasegawa Y, Saito T, Ogihara T, et al. Blockade of the
macrophage-gene expression in lipid metabolism and inflam- nuclear factor-κB pathway in the endothelium prevents
mation. Nat Med. 2001;7(1):48-52. insulin resistance and prolongs life spans. Circulation.
95. Kennedy DJ, Kuchibhotla S, Westfall KM, Silverstein RL, 2012;125(9):1122-1133.
Morton RE, Febbraio M. A CD36-dependent pathway en- 114. Barinda AJ, Ikeda K, Nugroho DB, et al. Endothelial progeria
hances macrophage and adipose tissue inflammation and im- induces adipose tissue senescence and impairs insulin sensi-
pairs insulin signalling. Cardiovasc Res. 2011;89(3):604-613. tivity through senescence associated secretory phenotype. Nat
96. Silverstein RL, Febbraio M. CD36, a scavenger receptor in- Commun. 2020;11(1):481.
volved in immunity, metabolism, angiogenesis, and behavior. 115. Imanishi T, Hano T, Sawamura T, Nishio I. Oxidized
Sci Signal. 2009;2(72):re3. low-density lipoprotein induces endothelial progenitor cell sen-
97. Daryabor G, Kabelitz D, Kalantar K. An update on immune escence, leading to cellular dysfunction. Clin Exp Pharmacol
dysregulation in obesity-related insulin resistance. Scand J Physiol. 2004;31(7):407-413.
Immunol. 2019;89(4):e12747. 116. Crespo-Garcia S, Tsuruda PR, Dejda A, et al. Pathological
98. Liu W, Shen SM, Zhao XY, Chen GQ. Targeted genes and angiogenesis in retinopathy engages cellular senescence and
interacting proteins of hypoxia inducible factor-1. Int J is amenable to therapeutic elimination via BCL-xL inhibition.
Biochem Mol Biol. 2012;3(2):165-178. Cell Metab. 2021;33(4):818-832.e7.
99. Sun K, Halberg N, Khan M, Magalang UJ, Scherer PE. Selective 117. Rojas M, Lemtalsi T, Toque HA, et al. NOX2-induced activa-
inhibition of hypoxia-inducible factor 1α ameliorates adipose tion of arginase and diabetes-induced retinal endothelial cell
tissue dysfunction. Mol Cell Biol. 2013;33(5):904-917. senescence. Antioxidants (Basel). 2017;6(2):43.
100. McDonnell WJ, Koethe JR, Mallal SA, et al. High CD8 T-cell 118. Shosha E, Xu Z, Narayanan SP, et al. Mechanisms of diabetes-
receptor clonality and altered CDR3 properties are associated induced endothelial cell senescence: role of arginase 1. Int J
with elevated isolevuglandins in adipose tissue during diet- Mol Sci. 2018;19(4):1215.
induced obesity. Diabetes. 2018;67(11):2361-2376. 119. Mortuza R, Chen S, Feng B, Sen S, Chakrabarti S. High glucose
101. Palmer AK, Xu M, Zhu Y, et al. Targeting senescent cells al- induced alteration of SIRTs in endothelial cells causes rapid
leviates obesity-induced metabolic dysfunction. Aging Cell. aging in a p300 and FOXO regulated pathway. PLoS One.
2019;18(3):e12950. 2013;8(1):e54514.
102. Schafer MJ, White TA, Evans G, et al. Exercise prevents 120. Thounaojam MC, Jadeja RN, Warren M, et al. MicroRNA-34a
diet-induced cellular senescence in adipose tissue. Diabetes. (miR-34a) mediates retinal endothelial cell premature senes-
2016;65(6):1606-1615. cence through mitochondrial dysfunction and loss of antioxi-
103. Covarrubias AJ, Kale A, Perrone R, et al. Senescent cells pro- dant activities. Antioxidants (Basel). 2019;8(9):328.
mote tissue NAD+ decline during ageing via the activation of 121. Liu J, Chen S, Biswas S, et al. Glucose-induced oxida-
CD38+ macrophages. Nat Metab. 2020;2(11):1265-1283. tive stress and accelerated aging in endothelial cells are
12  Endocrinology, 2021, Vol. 162, No. 10

mediated by the depletion of mitochondrial SIRTs. Physiol Rep. 136. Nishimura S, Manabe I, Nagasaki M, et al. CD8+ effector T
2020;8(3):e14331. cells contribute to macrophage recruitment and adipose tissue
122. Liu R, Liu H, Ha Y, Tilton RG, Zhang W. Oxidative stress in- inflammation in obesity. Nat Med. 2009;15(8):914-920.
duces endothelial cell senescence via downregulation of Sirt6. 137. Chuang HC, Sheu WH, Lin YT, et al. HGK/MAP4K4 defi-
Biomed Res Int. 2014;2014:902842. ciency induces TRAF2 stabilization and Th17 differentiation
123. Zhang T, Tian F, Wang J, et al. Donepezil attenuates high leading to insulin resistance. Nat Commun. 2014;5:4602.
glucose-accelerated senescence in human umbilical vein endo- 138. Surendar J, Frohberger SJ, Karunakaran I, et al. Adiponectin
thelial cells through SIRT1 activation. Cell Stress Chaperones. limits IFN-γ and IL-17 producing CD4 T cells in obesity
2015;20(5):787-792. by restraining cell intrinsic glycolysis. Front Immunol.
124. Binet F, Cagnone G, Crespo-Garcia S, et al. Neutrophil extra- 2019;10:2555.
cellular traps target senescent vasculature for tissue remodeling 139. Zhao RX, He Q, Sha S, et al. Increased AHR transcripts cor-
in retinopathy. Science. 2020;369(6506):eaay5356. relate with pro-inflammatory T-helper lymphocytes polariza-
125. Liu Y, Drozdov I, Shroff R, Beltran LE, Shanahan CM. tion in both metabolically healthy obesity and type 2 diabetic

Downloaded from https://academic.oup.com/endo/article/162/10/bqab136/6345039 by guest on 24 March 2024


Prelamin A accelerates vascular calcification via activation patients. Front Immunol. 2020;11:1644.
of the DNA damage response and senescence-associated se- 140. Bapat SP, Myoung Suh J, Fang S, et al. Depletion of fat-resident
cretory phenotype in vascular smooth muscle cells. Circ Res. Treg cells prevents age-associated insulin resistance. Nature.
2013;112(10):e99-109. 2015;528(7580):137-141.
126. Parvizi M, Ryan ZC, Ebtehaj S, Arendt BK, Lanza IR. The 141. Cipolletta D, Feuerer M, Li A, et al. PPAR-γ is a major driver
secretome of senescent preadipocytes influences the phenotype of the accumulation and phenotype of adipose tissue Treg cells.
and function of cells of the vascular wall. Biochim Biophys Nature. 2012;486(7404):549-553.
Acta Mol Basis Dis. 2021;1867(1):165983. 142. Cucak H, Vistisen D, Witte D, Philipsen A, Rosendahl A.
127. Cui XJ, Lin X, Zhong JY, et al. Adiponectin attenuates the pre- Reduction of specific circulating lymphocyte populations with
mature senescence of vascular smooth muscle cells induced by metabolic risk factors in patients at risk to develop type 2 dia-
high glucose through mTOR signaling pathway. Aging Med betes. PLoS One. 2014;9(9):e107140.
(Milton). 2020;3(3):178-187. 143. Feuerer M, Herrero L, Cipolletta D, et al. Lean, but not
128. Alicic RZ, Rooney MT, Tuttle KR. Diabetic kidney disease: obese, fat is enriched for a unique population of regula-
challenges, progress, and possibilities. Clin J Am Soc Nephrol. tory T cells that affect metabolic parameters. Nat Med.
2017;12(12):2032-2045. 2009;15(8):930-939.
129. Xiong Y, Zhou L. The signaling of cellular senescence in diabetic 144. Sharma M, Boytard L, Hadi T, et al. Enhanced glycolysis and
nephropathy. Oxid Med Cell Longev. 2019;2019:7495629. HIF-1α activation in adipose tissue macrophages sustains local
130. Zhang X, Chen X, Wu D, et al. Downregulation of connexin 43 and systemic interleukin-1β production in obesity. Sci Rep.
expression by high glucose induces senescence in glomerular 2020;10(1):5555.
mesangial cells. J Am Soc Nephrol. 2006;17(6):1532-1542. 145. Bertola A, Ciucci T, Rousseau D, et al. Identification of adipose
131. Chen K, Dai H, Yuan J, et al. Optineurin-mediated tissue dendritic cells correlated with obesity-associated insulin-
mitophagy protects renal tubular epithelial cells against ac- resistance and inducing Th17 responses in mice and patients.
celerated senescence in diabetic nephropathy. Cell Death Dis. Diabetes. 2012;61(9):2238-2247.
2018;9(2):105. 146. Coombes JL, Siddiqui KR, Arancibia-Cárcamo CV, et al.
132. Fyhrquist F, Tiitu A, Saijonmaa O, Forsblom C, Groop PH; A functionally specialized population of mucosal CD103+
FinnDiane Study Group. Telomere length and progression of DCs induces Foxp3+ regulatory T cells via a TGF-beta
diabetic nephropathy in patients with type 1 diabetes. J Intern and retinoic acid-dependent mechanism. J Exp Med.
Med. 2010;267(3):278-286. 2007;204(8):1757-1764.
133. Sampson MJ, Hughes DA. Chromosomal telomere attrition 147. Galle-Treger L, Sankaranarayanan I, Hurrell BP, et al.
as a mechanism for the increased risk of epithelial cancers Costimulation of type-2 innate lymphoid cells by GITR pro-
and senescent phenotypes in type 2 diabetes. Diabetologia. motes effector function and ameliorates type 2 diabetes. Nat
2006;49(8):1726-1731. Commun. 2019;10(1):713.
134. Kitada K, Nakano D, Ohsaki H, et al. Hyperglycemia causes 148. Shafiei-Jahani P, Hurrell BP, Galle-Treger L, et al. DR3 stimu-
cellular senescence via a SGLT2- and p21-dependent pathway lation of adipose resident ILC2s ameliorates type 2 diabetes
in proximal tubules in the early stage of diabetic nephropathy. mellitus. Nat Commun. 2020;11(1):4718.
J Diabetes Complications. 2014;28(5):604-611. 149. Zhao XY, Zhou L, Chen Z, et al. The obesity-induced adipokine
135. Guo J, Zheng HJ, Zhang W, et al. Accelerated Kidney Aging in sST2 exacerbates adipose Treg and ILC2 depletion and pro-
Diabetes Mellitus. Oxid Med Cell Longev. 2020;2020:1234059. motes insulin resistance. Sci Adv. 2020;6(20):eaay6191.

You might also like