Next Article in Journal
The Enhancement of Antimicrobial Photodynamic Therapy of Escherichia Coli by a Functionalized Combination of Photosensitizers: In Vitro Examination of Single Cells by Quantitative Phase Imaging
Next Article in Special Issue
Modeling of the Senescence-Associated Phenotype in Human Skin Fibroblasts
Previous Article in Journal
Mechanical Properties of TC11 Titanium Alloy and Graphene Nanoplatelets/TC11 Composites Prepared by Selective Laser Melting
Previous Article in Special Issue
Optimized Protocol for Proportionate CNS Cell Retrieval as a Versatile Platform for Cellular and Molecular Phenomapping in Aging and Neurodegeneration
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Age-Related Changes in the Fibroblastic Differon of the Dermis: Role in Skin Aging

1
Human Stem Cells Institute, 119333 Moscow, Russia
2
Department of Pharmacology, Institute of Pharmacy, I. M. Sechenov First Moscow State Medical University (Sechenov University), 119991 Moscow, Russia
3
N. N. Blokhin National Medical Research Oncology Center, Ministry of Health of Russia, 115478 Moscow, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(11), 6135; https://doi.org/10.3390/ijms23116135
Submission received: 12 April 2022 / Revised: 27 May 2022 / Accepted: 27 May 2022 / Published: 30 May 2022
(This article belongs to the Special Issue Cellular Senescence in Physiological and Pathological Processes)

Abstract

:
Skin aging is a multi-factorial process that affects nearly every aspect of skin biology and function. The processes developing in the skin during aging are based on fundamental molecular mechanisms associated with fibroblasts, the main cellular population of the dermis. It has been revealed that the amount of fibroblasts decreases markedly with age and their functional activity is also reduced. This inevitably leads to a decrease in the regenerative abilities of the skin and the progression of its aging. In this review we consider the mechanisms underlying these processes, mainly the changes observed with age in the stem/progenitor cells that constitute the fibroblastic differon of the dermis and form their microenvironment (niches). These changes lead to the depletion of stem cells, which, in turn, leads to a decrease in the number of differentiated (mature) dermal fibroblasts responsible for the production of the dermal extracellular matrix and its remodeling. We also describe in detail DNA damages, their cellular and systemic consequences, molecular mechanisms of DNA damage response, and also the role of fibroblast senescence in skin aging.

1. Quantitative Changes in the Population of Dermal Fibroblasts

With age, a decrease in the number of cells is observed in the population of dermal fibroblasts (DFs), which has been confirmed in a number of scientific papers. Varani et al. (2006) examined skin biopsies of young people (18–29 years old) and older people (80 years and older) isolated from areas of the body protected from ultraviolet radiation (UV). It has been shown that the total number of DFs in the older group was reduced by an average of 35% [1]. A study by Solé-Boldo et al. (2020), which was conducted using single-cell RNA sequencing (scRNA-seq) of DFs obtained from the photoprotected skin of young and elderly people (transcriptomes of more than 5000 DFs were analyzed), confirmed a decrease in the number of these cells with age. At the same time, lowering their heterogeneity was also revealed [2]. The decrease in the number of DFs has a significant impact on the biology of the skin since fibroblasts are the leading cells of the dermis, which is the connective tissue basis of the skin. They are responsible for the production, organization, and renovation of the components of the extracellular matrix (ECM) that determines the morphofunctional characteristics of the connective tissue [3,4,5,6,7].
Thus, a reduction in the amount of DFs leads to skin homeostasis disturbance and is accompanied by significant changes in the skin micro- and macrostructure [8]. These processes in the DF populations have a selective pattern. In particular, a study by Mine et al. (2008), analyzing the characteristics of papillary and reticular fibroblast populations of the dermis, showed that, with age, the most prominent changes are observed in the fibroblast population of the papillary layer [9]. This is probably due to the predominance of stem/progenitor cells (SCs, which are tissue-specific mesenchymal cells, precursors of fibroblasts of the skin outside the hair follicles and mesenchymal stem cells, such as bone marrow stem cells and others that were not incorporated into the skin during embryonic development) in the composition of the papillary fibroblast populations in which changes primarily occur [10].

2. Age-Related Changes in the Dermal Stem Cell Population

The main feature of SCs is the ability to self-renew and differentiate. By doing so, SCs play a key role in maintaining the tissue homeostasis providing the skin (as well as the other body tissues) with differentiated mature descendant cells [11,12,13,14]. It has been revealed that there is a decrease in the quantity and quality of the SCs functions with age [15,16], which leads to depletion of the SCs pool and, thus, to a decrease in the regenerative abilities of body tissues, one of the main causes of tissue aging [15,17,18,19,20,21,22,23].

2.1. Depletion of the SCs Pool

The depletion of the SCs pool can be caused by the whole range of factors, the main of which are the following:
  • A decrease in the ability of SCs to self-renew and the lowered proliferative potential [12,15,24], in particular, are due to a reduction in the number of receptors for growth factors located on the SCs’ surface membranes [25]. As a result, there is a decrease in the number of SCs capable of responding to signals stimulating their proliferation [26].
  • Excessive SC proliferation causing the depletion of the cell niche [18,19]. It has been revealed that the SCs’ regenerative potential is limited by a certain number of cell divisions during the life of the organism [27]; therefore, the prolonged activation of a stress factor stimulating the proliferation of SCs inevitably leads to depletion of their pool or to aberrant differentiation of these cells [28,29].
  • Increased apoptosis [22,30].
  • Telomere shortening [18,31].
  • Cellular aging [22,24,31,32].
Damages in the SCs themselves (and their accumulation with age) affect their DNA, mitochondrial (Mt) and protein complexes [11,18,33,34], epigenetic regulation of the genome (histone modification, DNA methylation, etc.), and other cellular structures [13]. The results of numerous studies indicate that the changes associated with the depletion of the SCs pool may be associated not only with disorders in the SCs themselves, but also with disorders in their niches (microenvironment) from which the cells constantly receive signals ensuring their proper functioning [21,35,36,37,38]. Thus, it has been revealed that disturbance of the ECM’s mechanical properties with age—in particular, an increase in its rigidity—can impair the SCs’ functions [39,40,41].
Studies by Sun et al. (2011) and Shakouri-Motlagh et al. (2017) showed that cultivation of MSCs (mesenchymal stem cells) isolated from “old” tissues on a native substrate significantly increases the ability of these cells to proliferate and differentiate [39,40] and, therefore, enhances their expansion. Koester et al. (2021) investigated the fundamental mechanisms of SCs aging by determining the age-dependent changes in the chromatin landscape of SCs isolated from mouse hair follicles (HFSCs). The results of the study also confirmed the significance of the “mechanical” state of the HFSCs niche that, in particular, affects the availability of chromatin and influences the epigenetic status of the key genes responsible for SCs’ self-renewing and differentiation [41]. It has been revealed that the ECM stiffness (in this study, the basement membrane) observed with age leads to a decrease in the availability of HFSCs chromatin. According to the author’s opinion, the rigid matrix of the SCs niche promotes the development of mechanosensory stress, which induces epigenetic changes in the genome that cause an impairment of the SCs’ functions.
Thus, the depletion of the SCs pool is a collective consequence of the many damages observed in the cells and in their microenvironment. However, it mainly correlates with disorders in the genetic apparatus of the cell [19], which lead to instability of the genome and finally terminating the proliferative life of the cell [36].

2.2. Changes in the SC Genome

Disorders in the genetic apparatus of the cell occur under the influence of a number of factors, both internal and external. The internal ones include [14,18]:
  • Reactive oxygen species (ROS) produced by metabolic intermediates and dysfunctional Mt;
  • DNA replication and DNA repair errors;
  • Glycation end products;
  • Dysfunction and shortening of telomeres (the end regions of chromosomes that ensure stable cell replication and protect chromosomes from fusion);
  • Inflammation of the surrounding tissue.
As a result of these factors action, about 100,000 DNA damages can occur per day in each cell, including various types of nucleotide modifications and single- and double-stranded breaks [14,42]. The second group of external factors includes UV and X-rays, chemicals, tobacco smoke, alcohol, and viral infections [12,36,43,44].
For skin cells, the most significant is the effect of ROS that promote the induction of both chronological skin aging and photoaging by damaging the nucleus and other structures/organelles of the cell [44,45,46,47,48]. During photoaging, ROS (superoxide anion (O2), hydrogen peroxide (H2O2), etc.) are formed under the influence of UV light as a result of photochemical reactions. During chronological skin aging, ROS generation regularly occurs in DFs and keratinocytes from molecular oxygen in the process of aerobic respiration and cellular metabolism [49].
The decrease in reparative abilities of the SC genome observed with age also plays a significant role in its instability. This process leads both to the impossibility to repair the occurring damages and to the emergence of errors in reparation, which cause new genomic damages and progressive accumulation of mutations [18,50,51]. A study by Garagnani et al. (2021), using the whole genome sequencing of centenarians (with average age of 106.6 years), confirmed that longevity is connected to a certain genetic pattern associated with effective DNA repair mechanisms and a low level of somatic mutations. According to the author’s opinion, this is the key factor of cellular homeostasis [52].
Changes in the SC genome appear to be the result of the realization of a genetic program that must ensure the genomic stability and the dynamics of expression of various genes at different stages of ontogenesis [5,53]. It has been shown that the aging process in humans is controlled by evolutionarily conservative genetic signaling pathways [18]. Genetic and biochemical analysis of SCs in various tissues has revealed age-related changes in DNA methylation, histone acetylation, and methylation markers. The research results suggest that epigenetic regulation is also the evolutionarily conservative mechanism involved in SCs aging in the human tissues [20,54,55].

3. Cellular Aging or Senescence

The state of genetic instability accompanied by a violation of the transcription process of vital genes and destroyed cellular metabolism leads to cellular aging [56], which is phenotypically associated with a wide range of changes at the molecular and cellular levels, including genomic (double-stranded DNA breaks, telomere shortening, decreased efficiency of repair processes) and epigenomic changes, Mt dysfunction, impairment of proteostasis and functions of other cellular structures, and deregulation of signaling pathways [12,13,36,57,58] (Figure 1).
Thus, DNA damage causes a response from the cell (DNA damage response, DDR), which is generally aimed at preventing the proliferation of damaged cells that are potentially oncogenic [18,31,59]. DDR involves a coordinated cascade of autonomous cellular and non-cellular reactions, including various pathways of DNA repair, cell apoptosis, and control points of the cell cycle, and its irreversible stop, causing cellular aging or senescence (Lat. senex, “old”) that, apparently, makes the most significant contribution to depletion of the SC pool [19,22,24,31,32,36,44,48,60,61,62] (Figure 2).
At the molecular level, the process of cellular aging is associated with the following mechanisms. The persistent damage to nuclear DNA triggers a activation of DNA damage response (DDR) pathways, which activates the signaling pathways of aging effectors - inhibitors of cyclin-dependent kinase p53-p21 and/or p16INK4a [63,64] (Figure 2) [36,65,66]. The p53-p21 signaling pathway is activated first, which initiates a temporary stop of the cell cycle, thereby preventing the transmission of damaged genetic information [63,64,65,66]. Then, if the damage is minor and subject to repair (it should be noted that most DNA damages can be corrected in the course of reparation), the cell cycle is restarted. In the case when the DNA damage cannot be corrected, the cell undergoes apoptosis. At the same time, the p21 signaling pathway, by producing the chemokine CXCL14, activates the genes of immune “surveillance”, attracting macrophages that control cells exposed to stress. This provides the first line of defense against dysfunctional cells. If the damaged cells are restored within four days, the macrophages abandon their targets. Otherwise, macrophages mobilize cytotoxic T lymphocytes, which promote the elimination of damaged cells [67].
In the case of persistent DNA damages (and/or their accumulation), an irreversible arrest of the cell cycle occurs, and the cell acquires a senescent phenotype (Figure 2) [19,36,62,68]. It has been determined that in response to cellular stress, protein p21 changes the regulatory features of pRb transcription (pRb is a retinoblastoma protein, a suppressor of carcinogenesis) and leads to inhibition of the cell cycle genes [36,69], while the protein p16 INK4a plays the key role and regulates the long-term cell cycle arrest via the pRb-E2F signaling pathway (pRb is the main mediator supporting the irreversibility of the cell cycle) [36,69]. With age, this protein accumulates significantly in tissues [66,70,71]).
Cellular senescence was first described in the 1960s, when L. Hayflick and P.S. Moorhead (1961) established that human fibroblasts in culture undergo a certain number of divisions (about 50) before cell expansion is stopped [72]. This limit of cell divisions, called the Hayflick limit, is a result of progressive shortening of telomeres with each cell division. This is a physiological reaction aimed at preventing the accumulation of DNA damages [73]. When telomere shortening reaches a critical level, DDR occurs, finally leading to either cell apoptosis or irreversible cell cycle arrest. One of the main causes of telomere shortening is a decrease in the level/activity of enzyme telomerase, which is observed with age [74]. This type of cellular aging is called replicative aging [62].
Along with replicative aging, all human diploid cells, including DFs, can undergo premature aging unrelated to telomere shortening [62,65]. The premature cellular aging can be divided into two types depending on the DNA damage triggers: (1) oncogen-induced cellular aging, developing in response to oncogen activation; and (2) stress-induced cellular aging, resulting from oxidative stress under the influence of ultraviolet light, ionizing radiation, and ROS [44,48,62,65,75]. For example, it has been shown using a DFs culture (in vitro) that tobacco smoke components cause oxidative DNA damage, leading to premature cell aging and further to irreversible cell cycle arrest [18,76].
It should be noted that regardless of the trigger that induces cellular aging and regardless of the type of cellular aging, the irreversible stop of the cell cycle is general in the process of cellular aging [74,75], as well as the acquisition by cells of the senescent phenotype (senCs, senDFs) (Figure 3).
The damage of the cell nucleus inevitably leads to impairment of other cellular structures/organelles, which is accompanied by an increase in degenerative processes. Thus, the increasing instability of the nuclear genome observed with age is critical for the functioning of mitochondria (Mt), which are the key organelles in eukaryotic cells that play a central role in ATP production and cell bioenergetics [56,77,78,79,80,81]. Many proteins necessary for Mt activity are encoded by nuclear genes and Mt dysfunction occurs when nuclear genes are damaged [80,82]. In addition, when DDR is activated, a decrease in the amount of Mt occurs, as well as exhaustion of their membrane potential and impairment in the electron transport chain and the mitophagy process [83,84]. As a result of Mt changes, the production of ROS increases in cells, which promotes the progressive damage of proteins and lipids, as well as the accumulation of lipofuscin (an insoluble covalently crosslinked glycolipoprotein complex) [85], telomere dysfunction [86], and degradation of lysosomes. All these processes cause an even greater increase in ROS production. Thus, a vicious autocrine feedback circle is initiated that increases the ROS-mediated damage to nuclear DNA, which leads to the progression of SCs aging [44,45,87,88,89,90,91,92].
It should be noted that the accumulation of lipofuscin underlying age-related skin pigmentation disorders is also promoted by the decrease in autophagy observed with age, which is characteristic of senDFs [93]. According to research results, Mt dysfunction serves as an important factor inducing both chronological aging and photo-aging of the skin, possibly creating a common link between them [44,94,95]. In the case of both types of skin aging, the high level of mutations (deletion (decay) of 4977 base pairs) has been revealed in mitochondrial DNA of DFs [96].
In the process of cellular aging, the disruption of proteostasis is also observed, which is manifested by the accumulation of modified and improperly folded proteins and protein aggregates, which occurs due to a decrease in the ability of cellular proteasomes to eliminate the protein aggregates [97,98,99]. This leads to a decrease in cellular functions, as well as to oxidation of the cellular membrane lipids, which promotes a reduction in the efficiency of transmembrane transport and a disruption in the transmembrane signaling pathways [24].
The abovementioned studies confirm the conclusion that cellular aging is a complex process involving structural damage and impairment of the vital activity of almost all cell organelles. As a consequence, an autocrine vicious circle is formed that violates cell functioning.

4. Senescent Fibroblasts and “Paracrine” Skin Aging

Senescent DFs (senDFs) like other senescent cells (senCs) undergo morphological changes, promoting their resistance to apoptosis, which contributes to the accumulation of these cells in the dermis [62,100] and changes the pattern of gene expression. The resistance of senDFs to apoptosis may be caused by an increase in the levels of anti-apoptotic proteins such as BCL-2 and FOXO4 [101,102]. It may also be associated with the secretion of non-classical MHC HLA-E, which activates inhibitory receptor NKG2A on natural killer cells (NK cells) and CD8+ T lymphocytes. Activation of NKG2A prevents the elimination of senDFs [103]. The altered pattern of gene expression leads to reprogramming of senCs metabolism and secretion of a wide range of soluble and insoluble factors which are commonly referred to secretory phenotype associated with aging (SASP, Senescence-Associated Secretory Phenotype) [104,105]. SASP includes many pro-inflammatory cytokines/chemokines, growth factors, metalloproteinases (MMPs), and other soluble factors (Figure 4) [46,106]. Due to secretion of these factors, cells with SASP by a “paracrine” way promote the aging of neighboring SCs, inflammation of the tissue, and destruction/degradation of dermal ECM through the action of MMPs. In addition to the soluble SASP factors, extracellular vesicles (exosomes) produced by senCs are also important mediators of the “paracrine” aging. By means of their constituent microRNAs, specific proteins, nucleic acids, and other insoluble factors, senCs perform intercellular connections, in particular between senDFs and neighboring SCs, enhancing the aging processes in these cells (as has been shown both in vitro and in vivo) [107].
Basisty et al. (2020) published the senDFs SASP atlas, which contains a large database, including the soluble factors and exosomes produced by senDFs. The authors revealed that SASP is variable and its composition may change depending on the trigger factor that causes cell aging [108]. Waldera Lupa et al. (2015), by examining the DFs obtained from the skin of elderly people with chronological aging, revealed a specific secretome, including 70 proteins corresponding to the classic SASP and 21 unique proteins that, according to the authors, may reflect specific skin-aging processes [109].

4.1. Identification of Senescent Fibroblasts

Senescent DFs (like other senCs) are cells that do not have specific markers; so, they can be identified by means of a combination of characteristic features [76]. The following senDFs markers are most frequently used in vitro (they were initially studied in vitro during the induced DFs aging, both replicative and premature [70,110]):
  • Morphological changes: the increase in size and flattening of the shape [111];
  • Increased activity of senescence-associated lysosomal enzyme β-galactosidase (SA-b-gal) which is the “gold standard” for the identification of senCs both in vitro and in vivo (in tissue samples) [112];
  • Increased levels of inhibitors of the cell cycle (p16INK4a, p21CIP1, and p53) [31,76];
  • Visualization of cytoplasmic granularity under a light microscope: it indicates an increase in the number and size of lysosomes (while this does not mean an increase in the activity of these organelles since there is a marked decrease in the level of autophagy associated with lysosomes during aging) [88,93,113];
  • Accumulation of lipofuscin [113];
  • Increase in the frequency of γH2AX (a marker of double-stranded DNA breaks that occur during persistent DNA damage and DDR activation [76]);
  • Presence of telomere-associated foci of DNA damage (TAF) [114];
  • Decrease in the level of nuclear intermediate plate protein and epigenetic modulator of lamin B1 [115,116,117,118,119] (the level of lamin B1 decreases in vitro in senDFs regardless of the stress factor [117] and in vivo in DFs isolated from skin samples with signs of premature and chronological aging [120]);
  • Presence of senescence-associated heterochromatin foci (SAHF, special heterochromatin structures formed in the nuclei of senCs) [121];
  • Presence of DNA foci with chromatin changes that enhance cell aging (DNA-SCARS, DNA Segments with Chromatin Alterations Reinforcing Senescence) [69];
  • Presence of HMGB1 (a protein from the group of nuclear non-histone proteins; in senCs it leaves the nucleus and moves to the cytoplasm and ECM; a decrease in its level in the nucleus leads to a decrease in gene expression) [122];
  • High levels of proinflammatory cytokines in cells with SASP (in particular, IL-1α, IL-6, TNF, NF-κB, etc.), chemokines (CXCR2), metalloproteinases (MMP-3 and MMP-9), etc. [123,124,125];
  • Mt dysfunction [84].
The DFs aging markers detected in vitro were also identified in vivo in the study of people with chronological aging [126,127] and photoaging [128]. Among these markers, the most common markers are the following:
  • High level of SA-β-gal activity;
  • Change in the production of ECM components;
  • Increased level of cyclin-dependent kinases p21 and p16INK4a (among the other markers detected in vivo, it has the highest correlation with markers revealed in vitro [31]);
  • Depletion of lamin B1 [117];
  • Presence of SAHF (Senescent-Associated Heterochromatin Foci) [70,129];
  • Increased level of SASP proinflammatory cytokines [124];
  • Presence of telomere-associated foci of DNA damage (TAF) used as the quantitative marker of skin tissue aging in situ [114].
In addition, the high level of VEGF (vascular endothelial growth factor) can serve as one of the markers of senDFs in vivo, which is associated with increased vascular permeability, erythema, and the risk of skin cancer. It should be noted that the activation of VEGF production in senDFs does not depend on the type of cellular aging trigger [130]. In the scientific literature, much attention was paid to the identification of senDFs in the skin since there is no doubt about the key role of these cells in the induction and progression of skin-tissue aging [24,44,95,131,132]. According to Gorgoulis et al. (2019), for the most reliable identification of senCs in tissues (in situ), a “multi-marker approach” should be used, combining, for example, cytoplasmic markers (SA-β-gal and lipofuscin) with nuclear markers (p16INK4A and p21WAF1/Cip1), as well as the markers associated with SASP [46]. The most frequently used methods are the following: single cell analysis, enzyme immunoassay, visualizing flow cytometry, or mass cytometry [46].

4.2. The Role of senDFs in Skin Aging

It has been revealed that, with age, an accumulation of senDFs occurs in the skin, which, through the associated specific SASP and by paracrine mechanisms, contributes to depletion of the SC pool, destruction/disruption of the regenerative abilities of the tissue, and, therefore, the induction and progression of its aging [14,44,132,133,134,135,136,137].
It should be noted that senCs belong to a specific type of cells. The results of preclinical studies have shown that selective removal of senCs significantly improves tissue homeostasis, prolongs the life of animals, and improves life quality [101,138]. However, it has also been shown that removal of senCs from the wound delays the healing, leads to fibrosis, and impairs the formation of granulation tissue [139]. In addition, it has been established that senCs participate in the remodeling/plasticity of body tissues [101]. These effects are related to the fact that cellular aging is an evolutionary antagonistic pleiotropic process while senCs are involved in every stage of ontogenesis: from embryogenesis, where they participate in the development of organs/tissues, to a mature state, where they play a significant role in the repair of organs/tissues by preventing the proliferation of cells with damaged DNA, thereby lowering the risk of neoplastic transformation [100,101,140].
In young people, the immune system effectively eliminates senCs from the body. However, the functions of the immune system decrease with age and the accumulation of senescent cells is observed in tissues [17]. As a result, the destruction of tissues occurs due to a number of pathophysiological processes mediated by SASP not only at the level of tissues but also at the level of cell populations.
At the level of tissues, degradation of the ECM is observed, caused by overproduction of MMPs in senescent cells [73]. Mild chronic aseptic inflammation also develops due to the secretion of many pro-inflammatory factors by senescent cells [14], including cytokines (IL-1, IL-6, and IL-8) and tumor necrosis factor (TNF), which recruit macrophages, neutrophils, and T cells infiltrating the tissues [100,141,142]. Thus, for example, it has been shown that IL-1 is directly associated with “paracrine” aging in vivo through activation of the inflammasome complex (a multi-protein oligomeric complex in myeloid cells responsible for activation of the inflammatory response) [135]. As a result, the following events occur: (1) degradation of membrane receptors, signaling pathways, proteins, and other components of the ECM; (2) changes in functions of the SCs niches; (3) disruption of autophagy processes; and (4) activation of transcription factor NF-κB (transcription nuclear factor of activated B cells), promoting the progression of inflammation in tissues [143]. The tissue inflammation is also accompanied by high levels of proinflammatory cytokines (IL-6, IL-1β, TGF-b, and TNF-α) that disrupt the transmission of anabolic signals, which leads to a decrease in the sensitivity of tissues to nutrients [144]. This close connection between aging and the inflammatory processes has been called “inflammaging” (inflammation + aging) or inflammatory aging.
At the level of cell populations, there are changes that are primarily associated with the paracrine mechanism, causing the aging phenotype in SCs located in spatial proximity to senCs [134,137,144,145,146,147,148], the so-called passive senescence [145] or “aging of an outsider observer” [146]. At the same time, the increase in the level of ROS is observed in senDFs, which causes the increase in Mt dysfunction. Thus, a vicious circle is formed (for more information, see above) that contributes to an even greater increase in ROS production and promotes the damaging effect of ROS on cell structures/organelles, which is accompanied by the reduced activity of antioxidant enzymes that is observed in cells with age [148,149]. In addition, factor NF-κB is activated and the “axis” ROS–NF-κB is formed, inducing DDR in SCs adjacent to senCs (the so-called “witness cells”) [44,48,94,147], which leads to the subsequent cell apoptosis or aging. At the same time, AP-1 and NF-B-dependent signaling pathways are also activated in senDFs, which contribute to the progression of inflammation in tissues [150].
Summarizing the abovementioned data and taking into account the main negative effects of senCs occurring in the skin [44,147], one can conclude the following about senescent cells:
  • They do not proliferate, which leads to violation of the SCs’ self-renewing process and depletion of the SCs pool;
  • Cause aging of neighboring SCs;
  • Promote an increase in the level of ROS and cause Mt dysfunction;
  • Induce DNA damage and aging of “witness cells” through the paracrine mechanism and ROS overproduction;
  • Cause the chronic aseptic inflammation in tissues due to the effect of proinflammatory SASP factors secreted by senCs;
  • Enhance the ECM degradation in the dermis by producing the high MMPs level;
  • Disrupt cellular and tissue homeostasis.
Thus, the pathogenetic processes associated with the accumulation of dysfunctional senDFs in skin tissues with age lead to depletion of the SC pool and, therefore, to the decrease in regenerative abilities and the progression of aging processes in the tissues (Figure 5) [44,48,136,146,148]. The main research data confirming the role of senDFs in skin aging are the following [44,95,151]:
  • An increase in the level of p16INK4a-positive DFs in the dermis correlating with the formation of wrinkles and the appearance of typical signs of elastic fiber aging;
  • During chronological aging, DFs have a proteomic profile in situ identical to the senDFs profile;
  • An increase in the ROS level leads to an increase in the number of p16INK4a-positive DFs in the skin and correlates with the progression of skin aging;
  • The spread of senescence to neighboring DFs with the expression of characteristic markers of cellular aging was recorded during transplantation of human senDFs into the skin of young immunodeficient mice;
  • Organ cultures obtained on the basis of human senDFs have signs of aging typical for chronological aging of the skin, including impairment of epidermal morphogenesis;
  • DFs isolated from the skin of elderly people are characterized by a gene expression pattern similar to that of senDFs;
  • Studies using a model of perforin-deficient mice (characterized by reduced functions of NK cells) have demonstrated the suppressed ability of the immune system to eliminate senDFs which by accumulating in the dermis lead to structural changes and the progression of aging processes in the dermis;
  • The results of a clinical study conducted using the local application of rapamycin on the skin of elderly people (with chronological aging) showed a decrease in the level of p16INK4a-positive DFs, as well as a decrease in the number of fine wrinkles and an increase in the thickness and elasticity of the skin.
  • Rapamycin is an inhibitor of mTOR (protein regulating the cell cycle and participating in the aging of DFs through the regulation of SASP) suppressing the translation of membrane-bound cytokine IL-1a and thereby inhibiting the secretion of pro-inflammatory SASP factors induced by IL-1a.

5. Conclusions

The considered molecular and cellular mechanisms observed in the skin during aging allow us to conclude that changes both in the SCs themselves and in their microenvironment (niches) occurring with age cause depletion of the SC population underlying the fibroblastic differon, which, in turn, leads to a decrease in the number of differentiated (mature) DFs responsible for ECM production and remodeling and, thus, to a decrease in the regenerative ability of the skin and the irreversible progression of skin aging.

Funding

This work was supported by Russian Science Foundation (RSF) (project No. 20-15-00321).

Conflicts of Interest

The authors declare no conflict of interests.

References

  1. Varani, J.; Dame, M.K. Decreased collagen production in chronologically aged skin: Roles of age-dependent alteration in fibroblast function and defective mechanical stimulation. Am. J. Pathol. 2006, 168, 1861–1868. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Solé-Boldo, L.; Raddatz, G. Single-cell transcriptomes of the human skin reveal age-related loss of fibroblast priming. Commun. Biol. 2020, 3, 188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Sorrell, J.M.; Caplan, A.I. Fibroblasts—A diverse population at the center of it all. Int. Rev. Cell. Molec. Biol. 2009, 276, 161–214. [Google Scholar] [CrossRef]
  4. Tracy, L.E.; Minasian, R.A. Extracellular matrix and dermal fibroblast function in the healing wound. Adv. Wound Care 2016, 5, 119–136. [Google Scholar] [CrossRef] [PubMed]
  5. Xue, M.; Zhao, R. Dermal Fibroblast Heterogeneity and Its Contribution to the Skin Repair and Regeneration. Adv. Wound Care 2022, 11, 87–107. [Google Scholar] [CrossRef]
  6. Woodley, D.T. Distinct Fibroblasts in the Papillary and Reticular Dermis: Implications for Wound Healing. Dermatol. Clin. 2017, 35, 95–100. [Google Scholar] [CrossRef]
  7. Fisher, G.; Kang, S. Mechanism of photoaging and chronological skin aging. Arch. Dermatol. 2002, 138, 1462–1467. [Google Scholar] [CrossRef]
  8. Gunin, A.G.; Petrov, V.V. Age-related changes in angiogenesis in human dermis. Exp. Gerontol. 2014, 55, 143–151. [Google Scholar] [CrossRef]
  9. Mine, S.; Fortunel, N.O. Aging Alters Functionally Human Dermal Papillary Fibroblasts but Not Reticular Fibroblasts: A New View of Skin Morphogenesis and Aging. PLoS ONE 2008, 3, 1–13. [Google Scholar] [CrossRef] [Green Version]
  10. Driskell, R.R.; Watt, F.M. Understanding fibroblast heterogeneity in the skin. Trends Cell Biol. 2015, 25, 92–99. [Google Scholar] [CrossRef]
  11. Goodell, M.A.; Rando, T.A. Stem cells and healthy aging. Science 2015, 350, 1199–1204. [Google Scholar] [CrossRef] [PubMed]
  12. Ren, R.; Ocampo, A.; Liu, G.H.; Belmonte, J.C.I. Regulation of stem cell aging by metabolism and epigenetics. Cell. Metab. 2017, 26, 460–474. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Sameri, S.; Samadi, P. Stem cell aging in lifespan and disease: A state-of-the-art review. Curr. Stem Cell Res. Ther. 2020, 15, 362–378. [Google Scholar] [CrossRef] [PubMed]
  14. Cinat, D.; Coppes, R.P. DNA Damage-Induced Inflammatory Microenvironment and Adult Stem Cell Response. Front. Cell. Dev. Biol. 2021, 9, 729136. [Google Scholar] [CrossRef] [PubMed]
  15. Jung, Y.; Brack, A.S. Cellular mechanisms of somatic stem cell aging. Curr. Topics Dev. Biol. 2014, 107, 405–438. [Google Scholar] [CrossRef] [Green Version]
  16. Stolzing, A.; Jones, E. Age-related changes in human bone marrow-derived mesenchymal stem cells: Consequences for cell therapies. Mech. Ageing Dev. 2008, 129, 163–173. [Google Scholar] [CrossRef] [PubMed]
  17. McHugh, D.; Gil, J. Senescence and aging: Causes, consequences, and therapeutic avenues. Cell. Biol. 2018, 217, 65–77. [Google Scholar] [CrossRef]
  18. Lopez-Otin, C.; Blasco, M.A. The hallmarks of aging. Cell 2013, 153, 1194–1217. [Google Scholar] [CrossRef] [Green Version]
  19. Rossi, D.J.; Jamieson, C.H. Stems cells and the pathways to aging and cancer. Cell 2008, 132, 681–696. [Google Scholar] [CrossRef] [Green Version]
  20. Chen, D.; Kerr, C. The Epigenetics of Stem Cell Aging Comes of Age. Trends Cell Biol. 2019, 29, 563–568. [Google Scholar] [CrossRef]
  21. Oh, J.; Lee, Y.D. Stem cell aging: Mechanisms, regulators and therapeutic opportunities. Nat. Med. 2014, 20, 870–880. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Wong, T.Y.; Solis, M.A. Molecular mechanism of extrinsic factors affecting antiaging of stem cells. World J. Stem Cells 2015, 7, 512–520. [Google Scholar] [CrossRef] [PubMed]
  23. Tümpel, S.; Rudolph, K.L. Quiescence: Good and Bad of Stem Cell Aging. Trends Cell Biol. 2019, 29, 672–685. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Jenkins, G. Molecular mechanisms of skin ageing. Mechanisms of Ageing and Development. Mech. Ageing Dev. 2002, 123, 801–810. [Google Scholar] [CrossRef]
  25. Wang, J.; Nirmala, X. Perspectives on Improving the Efficacy of PRP Treatment for Tendinopathy. J. Musculoskelet. Disord. Treat. 2016, 2, 1–5. [Google Scholar] [CrossRef]
  26. Zouboulis, C.; Adjaye, J. Human skin stem cells and the ageing process. Exp. Gerontol. 2008, 43, 986–997. [Google Scholar] [CrossRef] [Green Version]
  27. Pazhanisamy, S.K. Stem cells, DNA damage, ageing and cancer. Hematol. Oncol. Stem Cell Ther. 2009, 2, 375–384. [Google Scholar] [CrossRef] [Green Version]
  28. Barazzuol, L.; Ju, L. A coordinated DNA damage response promotes adult quiescent neural stem cell activation. PLoS Biol. 2017, 15, e2001264. [Google Scholar] [CrossRef] [Green Version]
  29. Tiwari, V.; Wilson, D.M., III. DNA damage and associated DNA repair defects in disease and premature aging. Am. J. Hum. Genet. 2019, 105, 237–257. [Google Scholar] [CrossRef] [Green Version]
  30. Kim, D.E.; Dollé, M.E.T. Deficiency in the DNA repair protein ERCC1 triggers a link between senescence and apoptosis in human fibroblasts and mouse skin. Aging Cell 2020, 19, e13072. [Google Scholar] [CrossRef]
  31. Sharpless, N.E.; DePinho, R.A. How stem cells age and why this makes us grow old. Nat. Rev. Mol. Cell Biol. 2007, 8, 703–713. [Google Scholar] [CrossRef] [PubMed]
  32. Cristofalo, V.J.; Pignolo, R.J. Replicative senescence of human fibroblast-like cells in culture. Physiol. Rev. 1993, 73, 617–638. [Google Scholar] [CrossRef] [PubMed]
  33. Price, F.D.; von Maltzahn, J. Inhibition of JAK-STAT signaling stimulates adult satellite cell function. Nat. Med. 2014, 20, 1174–1181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Sun, D.; Luo, M. Epigenomic profiling of young and aged HSCs reveals concerted changes during aging that reinforce self-renewal. Cell Stem Cell 2014, 14, 673–688. [Google Scholar] [CrossRef] [Green Version]
  35. Jones, D.L. No place like home: Anatomy and function of the stem cell niche. Nat. Rev. Mol. Cell Biol. 2008, 9, 11–21. [Google Scholar] [CrossRef]
  36. Schumacher, B.; Pothof, J. The central role of DNA damage in the ageing process. Nature 2021, 592, 695–703. [Google Scholar] [CrossRef]
  37. Morrison, S.J.; Spradling, A.C. Stem cells and niches: Mechanisms that promote stem cell maintenance throughout life. Cell 2008, 132, 598–611. [Google Scholar] [CrossRef] [Green Version]
  38. Scadden, D.T. Nice neighborhood: Emerging concepts of the stem cell niche. Cell 2014, 157, 41–50. [Google Scholar] [CrossRef] [Green Version]
  39. Sun, Y.; Li, W. Rescuing replication and osteogenesis of aged mesenchymal stem cells by exposure to a young extracellular matrix. FASEB J. 2011, 25, 1474–1485. [Google Scholar] [CrossRef] [Green Version]
  40. Shakouri-Motlagh, A.; O’Connor, A.J. Native and solubilized decellularized extracellular matrix: A critical assessment of their potential for improving the expansion of mesenchymal stem cells. Acta Biomater. 2017, 55, 1–12. [Google Scholar] [CrossRef]
  41. Koester, J.; Miroshnikova, Y.A. Niche stiffening compromises hair follicle stem cell potential during ageing by reducing bivalent promoter accessibility. Nat. Cell Biol. 2021, 23, 771–781. [Google Scholar] [CrossRef] [PubMed]
  42. Madabhushi, R.; Pan, L. DNA damage and its links to neurodegeneration. Neuron 2014, 83, 266–282. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Moskalev, A.A.; Shaposhnikov, M.V. The role of DNA damage and repair in aging through the prism of Koch-like criteria. Ageing Res. Rev. 2013, 12, 661–684. [Google Scholar] [CrossRef] [PubMed]
  44. Wlaschek, M.; Maity, P. Connective tissue and fibroblast senescence in skin aging. J. Investig. Dermatol. 2021, 141, 985–992. [Google Scholar] [CrossRef]
  45. Rinnerthaler, M.; Bischof, J. Oxidative stress in aging human skin. Biomolecules 2015, 5, 545–589. [Google Scholar] [CrossRef] [Green Version]
  46. Gorgoulis, V.; Adams, P.D. Cellular Senescence: Defining a Path Forward. Cell 2019, 179, 814–827. [Google Scholar] [CrossRef]
  47. Cavinato, M. Molecular mechanisms of UVB-induced senescence of dermal fibroblasts and its relevance for photoaging of the human skin. Exp. Gerontol. 2017, 94, 78–82. [Google Scholar] [CrossRef]
  48. Lee, Y.; Choi, S. Cellular Senescence and Inflammaging in the Skin Microenvironment. Int. J. Mol. Sci. 2021, 22, 3849. [Google Scholar] [CrossRef]
  49. Poljsak, B.; Milisav, I. The neglected significance of “antioxidative stress”. Oxidative Med. Cell. Longev. 2012, 2012, 480895. [Google Scholar] [CrossRef] [Green Version]
  50. Moskalev, A.A.; Proshkina, E.N. Genetics of Aging and Longevity. Vavilov J. Genet. Breed. 2016, 20, 426–440. [Google Scholar] [CrossRef] [Green Version]
  51. Beerman, I.; Seita, J. Quiescent hematopoietic stem cells accumulate DNA damage during aging that is repaired upon entry into cell cycle. Cell Stem Cell 2014, 15, 37–50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Garagnani, P.; Marquis, J. Whole-genome sequencing analysis of semi-supercentenarians. eLife 2021, 10, e57849. [Google Scholar] [CrossRef] [PubMed]
  53. Stephens, P.; Genever, P. Non-epithelial oral mucosal progenitor cell populations. Oral Dis. 2007, 13, 1–10. [Google Scholar] [CrossRef] [PubMed]
  54. Schworer, S.; Becker, F. Epigenetic stress responses induce muscle stem-cell ageing by Hoxa9 developmental signals. Nature 2016, 540, 428–432. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Sousa-Victor, P.; Ayyaz, A. Piwi is required to limit exhaustion of aging somatic stem cells. Cell Rep. 2017, 20, 2527–2537. [Google Scholar] [CrossRef] [Green Version]
  56. Proshkina, E.N.; Solovev, I.A. Key molecular mechanisms of aging, biomarkers, and potential interventions. Mol. Biol. 2020, 54, 883–921. [Google Scholar] [CrossRef]
  57. Niedernhofer, L.J.; Gurcar, A.U. Nuclear genomic instability and aging. Annu. Rev. Biochem. 2018, 87, 295–322. [Google Scholar] [CrossRef]
  58. Ben-Avraham, D. Epigenetics of aging. In Longevity Genes; Atzmon, G., Ed.; Springer: New York, NY, USA, 2015; Volume 847, pp. 179–191. [Google Scholar] [CrossRef]
  59. Collado, M.; Serrano, M. Senescence in tumours: Evidence from mice and humans. Nat. Rev. Cancer 2010, 10, 51–57. [Google Scholar] [CrossRef] [Green Version]
  60. Jackson, S.P.; Bartek, J. The DNA-damage response in human biology and disease. Nature 2009, 461, 1071–1078. [Google Scholar] [CrossRef] [Green Version]
  61. Scully, R.; Panday, A. DNA double-strand break repair-pathway choice in somatic mammalian cells. Nat. Rev. Mol. Cell Biol. 2019, 20, 698–714. [Google Scholar] [CrossRef]
  62. Calcinotto, A.; Kohli, J. Cellular senescence: Aging, cancer, and injury. Physiol. Rev. 2019, 99, 1047–1078. [Google Scholar] [CrossRef] [PubMed]
  63. Beausejour, C.M.; Krtolica, A. Reversal of human cellular senescence: Roles of the p53 and p16 pathways. EMBO J. 2003, 22, 4212–4222. [Google Scholar] [CrossRef] [PubMed]
  64. Schafera, M.J.; Millera, J.D. Cellular senescence: Implications for metabolic disease. Mol. Cell Endocrinol. 2017, 455, 93–102. [Google Scholar] [CrossRef]
  65. D’Adda di Fagagna, F. Living on a break: Cellular senescence as a DNA-damage response. Nat. Rev. Cancer 2008, 8, 512–522. [Google Scholar] [CrossRef] [PubMed]
  66. Hetz, C.; Chevet, E. Targeting the unfolded protein response in disease. Nat. Rev. Drug Discov. 2013, 12, 703–719. [Google Scholar] [CrossRef]
  67. Sturmlechner, I.; Zhang, C. p21 produces a bioactive secretome that places stressed cells under immunosurveillance. Science 2021, 374, eabb3420. [Google Scholar] [CrossRef]
  68. Campisi, J.; d’Adda di Fagagna, F. Cellular senescence: When bad things happen to good cells. Nat. Rev. Mol. Cell Biol. 2007, 8, 729–740. [Google Scholar] [CrossRef]
  69. Narita, M.; Nũnez, S. Rb-mediated heterochromatin formation and silencing of E2F target genes during cellular senescence. Cell 2003, 113, 703–716. [Google Scholar] [CrossRef] [Green Version]
  70. Waaijer, M.; Parish, W.E. The number of p16INK4a positive cells in human skin reflects biological age. Aging Cell 2012, 11, 722–725. [Google Scholar] [CrossRef]
  71. Krishnamurthy, J.; Torrice, C. Ink4a/Arf expression is a biomarker of aging. J. Clin Investig. 2004, 114, 1299–1307. [Google Scholar] [CrossRef]
  72. Hayflick, L.; Moorhead, P.S. The serial cultivation of human diploid cell strains. Exp. Cell Res. 1961, 25, 585–621. [Google Scholar] [CrossRef]
  73. Wiggins, K.A.; Clarke, M.C. Senescence utilises inflammatory caspases to drive the SASP. Aging 2019, 11, 3891–3892. [Google Scholar] [CrossRef] [PubMed]
  74. Maestroni, L.; Matmati, S. Solving the telomere replication problem. Genes 2017, 8, 55. [Google Scholar] [CrossRef] [PubMed]
  75. Zhang, H. Molecular signaling and genetic pathways of senescence: Its role in tumorigenesis and aging. J. Cell Physiol. 2007, 210, 567–574. [Google Scholar] [CrossRef] [PubMed]
  76. Wang, A.S.; Dreesen, O. Biomarkers of cellular senescence and skin aging. Front. Genet. 2018, 9, 247. [Google Scholar] [CrossRef]
  77. Detmer, S.A.; Chan, D.C. Functions and dysfunctions of mitochondrial dynamics. Nat. Rev. Mol. Cell Biol. 2007, 8, 870–879. [Google Scholar] [CrossRef] [PubMed]
  78. Moehle, E.A.; Shen, K. Mitochondrial proteostasis in the context of cellular and organismal health and aging. J. Biol. Chem. 2019, 294, 5396–5407. [Google Scholar] [CrossRef] [Green Version]
  79. Boyman, L.; Karbowski, M. Regulation of Mitochondrial ATP Production: Ca2+ Signaling and Quality Control. Trends Mol Med. 2020, 26, 21–39. [Google Scholar] [CrossRef]
  80. Annesley, S.J.; Fisher, P.R. Mitochondria in Health and Disease. Cells 2019, 8, 680. [Google Scholar] [CrossRef] [Green Version]
  81. Reynolds, J.C.; Bwiza, C.P. Mitonuclear genomics and aging. Hum. Genet. 2020, 139, 381–399. [Google Scholar] [CrossRef]
  82. Cao, K.; Feng, Z. Mitoepigenetics: An intriguing regulatory layer in aging and metabolic-related diseases. Free Radic. Biol. Med. 2021, 177, 337–346. [Google Scholar] [CrossRef] [PubMed]
  83. Balaban, R.; Nemoto, S. Mitochondria, oxidants, and aging. Cell 2005, 120, 483–495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Korolchuk, V.I.; Miwa, S. Mitochondria in cell senescence: Is mitophagy the weakest link? EBioMedicine 2017, 21, 7–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Gorgoulis, V.G.; Pefani, D.E. Integrating the DNA damage and protein stress responses during cancer development and treatment. J. Pathol. 2018, 246, 12–40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Koliada, A.K.; Krasnenkov, D.S. Telomeric aging: Mitotic clock or stress indicator? Front. Genet. 2015, 6, 82. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Park, C.B.; Larsson, N.G. Mitochondrial DNA mutations in disease and aging. J. Cell Biol. 2011, 193, 809–818. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Park, J.T. Adjustment of the lysosomal-mitochondrial axis for control of cellular senescence. Ageing Res. Rev. 2018, 47, 176–182. [Google Scholar] [CrossRef]
  89. Huang, H.; Manton, K.G. The role of oxidative damage in mitochondria during aging: A review. Front. Biosci. 2004, 9, 1100–1117. [Google Scholar] [CrossRef] [Green Version]
  90. Treiber, N.; Maity, P. Accelerated aging phenotype in mice with conditional deficiency for mitochondrial superoxide dismutase in the connective tissue. Aging Cell 2011, 10, 239–254. [Google Scholar] [CrossRef]
  91. Wiley, C.D.; Velarde, M.C. Mitochondrial dysfunction induces senescence with a distinct secretory phenotype. Cell Metab. 2016, 23, 303–314. [Google Scholar] [CrossRef] [Green Version]
  92. Gu, Y.; Han, J. Biomarkers, oxidative stress and autophagy in skin aging. Ageing Res. Rev. 2020, 59, 101036. [Google Scholar] [CrossRef]
  93. Ott, C.; König, J. Macroautophagy is impaired in old murine brain tissue as well as in senescent human fibroblasts. Redox Biol. 2016, 10, 266–273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Krutmann, J.; Schroeder, P. Role of mitochondria in photoaging of human skin: The defective powerhouse model. J. Investig. Dermatol. Symp. Proc. 2009, 14, 44–49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Fitsiou, E.; Pulido, T. Cellular senescence and the senescence-associated secretory phenotype as drivers of skin photoaging. J. Investig. Dermatol. 2020, 141, 1119–1126. [Google Scholar] [CrossRef] [PubMed]
  96. Chung, K.W.; Choi, Y.J. Molecular insights into SIRT1 protection against UVB-induced skin fibroblast senescence by suppression of oxidative stress and p53 acetylation. J. Gerontol A Biol. Sci. Med. Sci. 2015, 70, 959e68. [Google Scholar] [CrossRef] [PubMed]
  97. Castro, J.P.; Ott, C. Carbonylation of the cytoskeletal protein actin leads to aggregate formation. Free Radic. Biol. Med. 2012, 53, 916–925. [Google Scholar] [CrossRef]
  98. Grune, T. Oxidized protein aggregates: Formation and biological effects. Free Radic. Biol. Med. 2020, 150, 120–124. [Google Scholar] [CrossRef]
  99. Kaushik, S.; Cuervo, A.M. Proteostasis and aging. Nat. Med. 2015, 21, 1406–1415. [Google Scholar] [CrossRef]
  100. Munoz-Espin, D.; Serrano, M. Cellular senescence: From physiology to pathology. Nat. Rev. Mol. Cell Biol. 2014, 15, 482–496. [Google Scholar] [CrossRef]
  101. Yosef, R.; Pilpel, N. Directed elimination of senescent cells by inhibition of BCL-W and BCL-XL. Nat. Commun. 2016, 7, 11190. [Google Scholar] [CrossRef]
  102. Baar, M.P.; Brandt, R.M. Targeted apoptosis of senescent cells restores tissue homeostasis in response to chemotoxicity and aging. Cell 2017, 169, 132–147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Pereira, B.I.; Devine, O.P. Senescent cells evade immune clearance via HLA-E-mediated NK and CD8(+) T cell inhibition. Nat. Commun. 2019, 10, 2387. [Google Scholar] [CrossRef] [PubMed]
  104. Demaria, M.; Desprez, P. Cell Autonomous and Non-autonomous Effects of Senescent Cells in the Skin. J. Investig. Dermatol. 2015, 35, 1722–1726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Herranz, N.; Gil, J. Mechanisms and functions of cellular senescence. J. Clin. Investig. 2018, 128, 1238–1246. [Google Scholar] [CrossRef]
  106. Rodier, F.; Coppe, J.-P. Persistent DNA damage signalling triggers senescence-associated inflammatory cytokine secretion. Nat. Cell Biol. 2009, 11, 973–979. [Google Scholar] [CrossRef]
  107. Borghesan, M.; Fafián-Labora, J. Small Extracellular Vesicles Are Key Regulators of Non-cell Autonomous Intercellular Communication in Senescence via the Interferon Protein IFITM. Cell Rep. 2019, 27, 3956–3971. [Google Scholar] [CrossRef] [Green Version]
  108. Basisty, N.; Kale, A. A proteomic atlas of senescence-associated secretomes for aging biomarker development. PLoS Biol. 2020, 18, e3000599. [Google Scholar] [CrossRef] [Green Version]
  109. Waldera Lupa, D.M.; Kalfalah, F. Characterization of skin aging-associated secreted proteins (SAASP) produced by dermal fibroblasts isolated from intrinsically aged human skin. J. Investig. Dermatol. 2015, 135, 1954–1956. [Google Scholar] [CrossRef] [Green Version]
  110. Gruber, F.; Kremslehner, C. Cell aging and cellular senescence in skin aging—Recent advances in fibroblast and keratinocyte biology. Exp. Gerontol. 2020, 130, 110780. [Google Scholar] [CrossRef]
  111. Hayflick, L. The limited in vitro lifetime of human diploid cell strains. Exp. Cell Res. 1965, 37, 614–636. [Google Scholar] [CrossRef]
  112. Dimri, G.P.; Lee, X. A biomarker that identifies senescent human cells in culture and in aging skin in vivo. Proc. Natl. Acad. Sci. USA 1995, 92, 9363–9367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Eckhart, L.; Tschachler, E. Autophagic control of skin aging. Front. Cell Dev. Biol. 2019, 7, 143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Hewitt, G.; Jurk, D. Telomeres are favoured targets of a persistent DNA damage response in ageing and stress-induced senescence. Nat. Commun. 2012, 3, 708. [Google Scholar] [CrossRef] [PubMed]
  115. Waaijer, M.E.C.; Gunn, D.A. Do senescence markers correlate in vitro and in situ within individual human donors? Aging 2018, 10, 278–289. [Google Scholar] [CrossRef] [Green Version]
  116. Scaffidi, P.; Misteli, T. Lamin A-dependent nuclear defects in human aging. Science 2006, 312, 1059–1163. [Google Scholar] [CrossRef] [Green Version]
  117. Freund, A.; Laberge, R.-M. Lamin B1 loss is a senescence-associated biomarker. Mol. Biol. Cell 2012, 23, 2066–2075. [Google Scholar] [CrossRef]
  118. Taimen, P.; Pflegar, K. A progeria mutation reveals functions for lamin A in nuclear assembly, architecture, and chromosome organization. Proc. Natl. Acad. Sci. USA 2009, 106, 20788–20793. [Google Scholar] [CrossRef] [Green Version]
  119. Dreesen, O.; Chojnowski, A. Lamin B1 fluctuations have differential effects on cellular proliferation and senescence. J. Cell Biol. 2013, 200, 605–617. [Google Scholar] [CrossRef] [Green Version]
  120. Bernardes de Jesus, B.; Blasco, M.A. Assessing Cell and Organ Senescence Biomarkers. Circ. Res. 2012, 111, 97–109. [Google Scholar] [CrossRef] [Green Version]
  121. Sun, L.; Yu, R. Chromatin architectural changes during cellular senescence and aging. Genes 2018, 9, 211. [Google Scholar] [CrossRef] [Green Version]
  122. Davalos, A.R.; Kawahara, M. p53-dependent release of Alarmin HMGB1 is a central mediator of senescent phenotypes. J. Cell Biol. 2013, 201, 613–629. [Google Scholar] [CrossRef] [PubMed]
  123. Tilstra, J.S.; Robinson, A.R. NF-κB inhibition delays DNA damage-induced senescence and aging in mice. J. Clin. Investig. 2012, 122, 2601–2612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Coppe, J.P.; Desprez, P.Y. The senescence-associated secretory phenotype: The dark side of tumor suppression. Ann. Rev. Pathol. 2010, 5, 99–118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Acosta, J.C. Chemokine signaling via the CXCR2 receptor reinforces senescence. Cell 2008, 133, 1006–1018. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Ressler, S.; Bartkova, J. p16INK4A is a robust in vivo biomarker of cellular aging in human skin. Aging Cell 2006, 5, 379–389. [Google Scholar] [CrossRef]
  127. Campisi, J. Aging, Cellular Senescence, and Cancer. Ann. Rev. Physiol. 2013, 75, 685–705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Debacq-Chainiaux, F.; Leduc, C. UV, stress and aging. Dermato-Endocrinol. 2012, 4, 236–240. [Google Scholar] [CrossRef] [Green Version]
  129. Aird, K.M.; Zhang, R. Detection of senescence-associated heterochromatin foci (SAHF). Methods Mol. Biol. 2013, 965, 185–196. [Google Scholar] [CrossRef] [Green Version]
  130. Coppe, J.P.; Kauser, K. Secretion of vascular endothelial growth factor by primary human fibroblasts at senescence. J. Biol. Chem. 2006, 281, 29568e74. [Google Scholar] [CrossRef] [Green Version]
  131. Velarde, M.C.; Demaria, M. Targeting senescent cells: Possible implications for delaying skin aging: A mini-review. Gerontology 2016, 62, 513–518. [Google Scholar] [CrossRef] [Green Version]
  132. Ho, C.-Y.; Dreesen, O. Faces of cellular senescence in skin aging. Mech. Age Dev. 2021, 198, 111525. [Google Scholar] [CrossRef] [PubMed]
  133. Childs, B.G.; Durik, M. Cellular senescence in aging and age-related disease: From mechanisms to therapy. Nat. Med. 2015, 21, 1424–1435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Da Silva, P.F.L.; Schumacher, B. DNA damage responses in ageing. Open Biol. 2019, 9, 190168. [Google Scholar] [CrossRef] [PubMed]
  135. Acosta, J.C.; Banito, A. A complex secretory program orchestrated by the inflammasome controls paracrine senescence. Nat. Cell Biol. 2013, 15, 978–990. [Google Scholar] [CrossRef] [PubMed]
  136. Tchkonia, T.; Zhu, Y. Cellular senescence and the senescent secretory phenotype: Therapeutic opportunities. J. Clin. Investig. 2013, 123, 966–972. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Toutfaire, M.; Bauwens, E. The impact of cellular senescence in skin ageing: A notion of mosaic and therapeutic strategies. Biochem. Pharmacol. 2017, 142, 1–12. [Google Scholar] [CrossRef]
  138. Baker, D.J.; Childs, B.G. Naturally occurring p16 Ink4a-positive cells shorten healthy lifespan. Nature 2016, 530, 184e9. [Google Scholar] [CrossRef] [Green Version]
  139. Demaria, M.; Ohtani, N. An essential role for senescent cells in optimal wound healing through secretion of PDGF-AA. Dev. Cell 2014, 31, 722. [Google Scholar] [CrossRef] [Green Version]
  140. Muñoz-Espín, D.; Cañamero, M. Programmed cell senescence during mammalian embryonic development. Cell 2013, 155, 1104–1118. [Google Scholar] [CrossRef] [Green Version]
  141. Fafián-Labora, J.A.; O’Loghlen, A. NF-κB/ IKK activation by small extracellular vesicles within the SASP. Aging Cell 2021, 20, e13426. [Google Scholar] [CrossRef]
  142. Kale, A.; Sharma, A. Role of immune cells in the removal of deleterious senescent cells. Immun. Ageing 2020, 17, 16. [Google Scholar] [CrossRef] [PubMed]
  143. Salminen, A.; Kaarniranta, K. Inflammaging: Disturbed interplay between autophagy and inflammasomes. Aging 2012, 4, 166–175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Van Deursen, J.M. The role of senescent cells in ageing. Nature 2014, 509, 439–446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Nelson, G.; Wordsworth, J. A senescent cell bystander effect: Senescence-induced senescence. Aging Cell 2012, 11, 345–349. [Google Scholar] [CrossRef] [Green Version]
  146. Da Silva, P.F.L.; Ogrodnik, M. The bystander effect contributes to the accumulation of senescent cells in vivo. Aging Cell 2019, 18, e12848. [Google Scholar] [CrossRef]
  147. Nelson, G.; Kucheryavenko, O. The senescent bystander effect is caused by ROS-activated NF-kB signaling. Mech. Ageing Dev. 2018, 170, 30–36. [Google Scholar] [CrossRef]
  148. Faget, D.V.; Ren, Q. Unmasking senescence: Context-dependent effects of SASP in cancer. Nat. Rev. Cancer 2019, 19, 439–453. [Google Scholar] [CrossRef]
  149. Tomaru, U.; Takahashi, S. Decreased proteasomal activity causes age-related phenotypes and promotes the development of metabolic abnormalities. Am. J. Pathol. 2012, 180, 963–972. [Google Scholar] [CrossRef]
  150. Meyer, P.; Maity, P. A model of the onset of the senescence associated secretory phenotype after DNA damage induced senescence. PLoS Comput. Biol. 2017, 13, e1005741. [Google Scholar] [CrossRef] [Green Version]
  151. Laberge, R.M.; Sun, Y. mTOR regulates the pro-tumorigenic senescence-associated secretory phenotype by promoting il1a translation. Nat. Cell Biol. 2015, 17, 1049–1061. [Google Scholar] [CrossRef]
Figure 1. Characteristics of the cellular and molecular signs of cellular senescence.
Figure 1. Characteristics of the cellular and molecular signs of cellular senescence.
Ijms 23 06135 g001
Figure 2. Consequences of DNA damage at the molecular, cellular, and tissue levels.
Figure 2. Consequences of DNA damage at the molecular, cellular, and tissue levels.
Ijms 23 06135 g002
Figure 3. Schematic representation of cellular aging.
Figure 3. Schematic representation of cellular aging.
Ijms 23 06135 g003
Figure 4. Components of the senescence-associated secretory phenotype (SASP).
Figure 4. Components of the senescence-associated secretory phenotype (SASP).
Ijms 23 06135 g004
Figure 5. Schematic representation of mechanisms of the tissue dysfunction involving senCs (SASP). SenCs accumulating with age in tissues contribute through the SASP factors to the depletion of the SC pool, degradation of ECM in the dermis, development of the chronic aseptic inflammation in tissues, and disruption of the tissue regenerative potential.
Figure 5. Schematic representation of mechanisms of the tissue dysfunction involving senCs (SASP). SenCs accumulating with age in tissues contribute through the SASP factors to the depletion of the SC pool, degradation of ECM in the dermis, development of the chronic aseptic inflammation in tissues, and disruption of the tissue regenerative potential.
Ijms 23 06135 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zorina, A.; Zorin, V.; Kudlay, D.; Kopnin, P. Age-Related Changes in the Fibroblastic Differon of the Dermis: Role in Skin Aging. Int. J. Mol. Sci. 2022, 23, 6135. https://doi.org/10.3390/ijms23116135

AMA Style

Zorina A, Zorin V, Kudlay D, Kopnin P. Age-Related Changes in the Fibroblastic Differon of the Dermis: Role in Skin Aging. International Journal of Molecular Sciences. 2022; 23(11):6135. https://doi.org/10.3390/ijms23116135

Chicago/Turabian Style

Zorina, Alla, Vadim Zorin, Dmitry Kudlay, and Pavel Kopnin. 2022. "Age-Related Changes in the Fibroblastic Differon of the Dermis: Role in Skin Aging" International Journal of Molecular Sciences 23, no. 11: 6135. https://doi.org/10.3390/ijms23116135

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop