Review Open Access
Copyright ©The Author(s) 2019. Published by Baishideng Publishing Group Inc. All rights reserved.
World J Stem Cells. Oct 26, 2019; 11(10): 787-802
Published online Oct 26, 2019. doi: 10.4252/wjsc.v11.i10.787
Aging: A cell source limiting factor in tissue engineering
Mohammadhossein Khorraminejad-Shirazi, Mohammadreza Dorvash, Alireza Estedlal, Amir Human Hoveidaei, Mohsen Mazloomrezaei, Pouria Mosaddeghi, Student Research Committee, Shiraz University of Medical Sciences, Shiraz 7134845794, Iran
Mohammadhossein Khorraminejad-Shirazi, Mohammadreza Dorvash, Alireza Estedlal, Mohsen Mazloomrezaei, Pouria Mosaddeghi, Cell and Molecular Medicine Student Research Group, School of Medicine, Shiraz University of Medical Sciences, Shiraz 7134845794, Iran
Mohammadreza Dorvash, Pouria Mosaddeghi, Pharmaceutical Sciences Research Center, Shiraz University of Medical Sciences, Shiraz 7134814336, Iran
ORCID number: Mohammadhossein Khorraminejad-Shirazi (0000-0001-8611-089X); Mohammadreza Dorvash (0000-0002-1935-9625); Alireza Estedlal (0000-0002-5317-3963); Amir Human Hoveidaei (0000-0003-4607-354X); Mohsen Mazloomrezaei (0000-0002-9296-9876); Pouria Mosaddeghi (0000-0003-0642-5248).
Author contributions: All authors contributed to drafting the manuscript and literature review; Estedlal A and Hoveidaei AH edited the manuscript; Khorraminejad-Shirazi M designed the concept of the study and provided administrative support; Khorraminejad-Shirazi M and Dorvash M did critical revision and final editing; Khorraminejad-Shirazi M and Dorvash M contributed equally to the manuscript; All authors have read and approved the final version of the manuscript.
Conflict-of-interest statement: No potential conflict of interest. No financial support.
Open-Access: This article is an open-access article which was selected by an in-house editor and fully peer-reviewed by external reviewers. It is distributed in accordance with the Creative Commons Attribution Non Commercial (CC BY-NC 4.0) license, which permits others to distribute, remix, adapt, build upon this work non-commercially, and license their derivative works on different terms, provided the original work is properly cited and the use is non-commercial. See: http://creativecommons.org/licenses/by-nc/4.0/
Corresponding author: Mohammadhossein Khorraminejad-Shirazi, MD, Department of Pathology, Cell and Molecular Medicine Student Research Group, Karim Khan Zand Boulevard, School of Medicine, Shiraz 7134845794, Iran. hkhorraminejad@sums.ac.ir
Telephone: +98-713-2083335
Received: February 18, 2019
Peer-review started: February 20, 2019
First decision: April 16, 2019
Revised: May 3, 2019
Accepted: September 4, 2019
Article in press: September 5, 2019
Published online: October 26, 2019

Abstract

Tissue engineering has yet to reach its ideal goal, i.e. creating profitable off-the-shelf tissues and organs, designing scaffolds and three-dimensional tissue architectures that can maintain the blood supply, proper biomaterial selection, and identifying the most efficient cell source for use in cell therapy and tissue engineering. These are still the major challenges in this field. Regarding the identification of the most appropriate cell source, aging as a factor that affects both somatic and stem cells and limits their function and applications is a preventable and, at least to some extents, a reversible phenomenon. Here, we reviewed different stem cell types, namely embryonic stem cells, adult stem cells, induced pluripotent stem cells, and genetically modified stem cells, as well as their sources, i.e. autologous, allogeneic, and xenogeneic sources. Afterward, we approached aging by discussing the functional decline of aged stem cells and different intrinsic and extrinsic factors that are involved in stem cell aging including replicative senescence and Hayflick limit, autophagy, epigenetic changes, miRNAs, mTOR and AMPK pathways, and the role of mitochondria in stem cell senescence. Finally, various interventions for rejuvenation and geroprotection of stem cells are discussed. These interventions can be applied in cell therapy and tissue engineering methods to conquer aging as a limiting factor, both in original cell source and in the in vitro proliferated cells.

Key Words: Aging, Senescence, Rejuvenation, Geroprotection, Tissue engineering, Stem cell therapy

Core tip: To attain profitable off-the-shelf tissues and organs, we must deal with the challenge of identifying and isolating an optimal cell source. Different types of stem cells with different properties have been used in tissue engineering and cell therapy to face this challenge. Although aging is an inevitable process that can eventually limit the function and stemness of stem cells, it is a conquerable phenomenon. In this article, we have reviewed several applicable interventions that can be used to overcome cellular aging.



INTRODUCTION

Anatomical and functional complexities of biological systems challenge the artificial construction of viable human tissues and organs. Proper three-dimensional tissue architecture to maintain blood supply is a key constraint on the size of the in vitro fabricated tissues[1]. In addition, biomaterial selection and strategies to design tissue scaffolds are vital for regulating cell signaling pathways that provide appropriate cell-cell interactions such as growth factor delivery, which is essential for cell differentiation. Although numerous attempts were made to overcome these key challenges in tissue engineering, the reproducible in vitro construction of artificial vascularized tissue is still needed[2].

The ideal goal of tissue engineering is to create off-the-shelf tissues and organs providing vast opportunities to tackle a group of diseases and to reduce the need for organ donors. This not only would treat millions of patients, but also increase human longevity and quality of life[1,3,4]. As the field of tissue engineering evolves, new obstacles appear in the way of the research and clinical application of these artificial tissues and organs. The fundamentals of this interdisciplinary field not only involves identifying biomaterials and designing scaffolds for in vivo cell expansion but also requires addressing the reliable cell sources. Hence, gradual advances in the clinical application of tissue engineering deal with hurdles in diverse aspects of science such as cell biology, bioengineering, and material science[5].

Apart from these engineering challenges, biologic issues and the major concern of identifying the ideal cell source is the other essential principle of tissue engineering[2]. Various stem cell types and sources have been extensively employed in regenerative medicine studies. However, each source has its own practical and technical challenges concerning their availability, isolation and cell expansion, cell delivery, aging, immunological barriers, and clinical and therapeutic efficiency. Furthermore, while major challenges of tissue engineering must be addressed at first, aging, as a cell source limiting factor, should not be overlooked. In this article, we have reviewed the cell sources that are used in tissue engineering and cell therapy techniques and how aging and cell senescence can challenge the isolation of ideal cell source. Also, we have discussed potentially applicable approaches for rejuvenation of aged cells.

CELL SOURCE AS A MAJOR CHALLENGE

First and foremost, the unresolved controversy of identifying the optimal cell types for tissue engineering is still a major challenge[4,6,7]. While cell transplantation, organ transplantation, and tissue engineering are fundamentally different, there are essentially three varieties of sources: Autologous, allogeneic, and xenogeneic cells, each of which can be subdivided into several types of stem cells including adult and embryonic stem cells. In addition, the discovery of induced pluripotent stem cells (iPSCs), which are discussed in the following sections, represent a promising source of cells for all branches of regenerative medicine[8,9].

Autologous sources

In autologous transplantation, the donor and the recipient are the same. Concerning the role of the immune system in potential tissue rejections, utilizing a patient’s own cells or “autologous cells” would be ideal. This method minimizes the chance of graft versus host disease and transmitted infections, and more importantly it would eliminate the need for lifetime use of immunosuppressive drugs, which improves the quality of life in post-transplant patients[10]. Despite these benefits, autologous cell therapy brings about several challenges. In fact, using the patient’s own cells might not be practical for the majority of cases. Transplant waiting lists are filled with aged patients who suffer from age-associated morbidities and cellular senescence affecting both their somatic and stem cells[11]. In addition, the patients who suffer from gene defects cannot easily benefit from autologous cell therapy[12]. Furthermore, to be viable for tissue engineering, millions of autologous cells should be collected from a donor and expanded ex vivo. For many tissue types, harvesting a sufficient number of cells is not applicable, especially when a patient is aged or severely diseased. Moreover, cell culture per se can cause undefined complications; the proliferative potential and clonogenicity of stem cells decrease after several cell divisions, which raises concerns about viability and functionality of cells after transplantation. These issues make autologous cell therapy undesirable for clinical applications, especially in emergencies or acute phases of disease[9,13].

Allogeneic sources

As mentioned earlier, the goal of tissue engineering is to manufacture large quantities of off-the-shelf tissues and organs that are immediately available to be administered clinically[14]. Allogeneic cells are cells from a genetically non-identical donor but of the same species. Thus, unaffected cells, tissues, and organs of every healthy donor can be a precious allogeneic cell source. This will rule out the challenges of aging, unavailability, and in vitro expansion challenges of autologous cell sources and consequently introduce allogeneic cell therapy as a promising method in case of emergency. This advantageousness paved the way for preparing a master bank of ready-made, clinically practical, and off-the-shelf allogeneic cells. On the contrary, the immunogenicity of allogeneic cells and the major histocompatibility complex (commonly known as MHC) incompatibilities are by far the most formidable barriers of allotransplantation. In addition, the side effects of immunosuppression like metabolic disorders, malignancies, and opportunistic infections can aggravate the outcome of a transplantation[9,12,15].

Xenogeneic sources

Xenogeneic or cross-species transplantation is the process of transplanting living cells, tissues, or organs from one species to another. In recent decades, the ever increasing demand for clinical transplantation and shortage of allogeneic sources for patients on the waiting lists has led to considerable amounts of clinical and experimental research in order to overcome the barriers of xenotransplantation. However, a great number of ethical red tape and immunological roadblocks are yet to be surpassed. Graft rejection and failure to achieve successful long-term outcomes are the main issues to be addressed, as there are great disparities between MHCs of different species. Another concern is the risk of zoonotic infections, particularly unidentified viruses. In addition, xenotransplantation is by itself a controversial ethical issue and sometimes raises religious concerns because it involves sacrificing animals to harvest organs and tissues for human usage[16].

CELL TYPES

Thus far, several stem cell types have been utilized in the field of tissue engineering.

Embryonic stem cells (ESCs)

ESCs are pluripotent stem cells isolated from the inner cell mass of blastocysts up until day 5.5 post-fertilization, right before the stage in which the embryo is ready for gastrulation[17,18]. They have unlimited potential for self-renewal and differentiation to be used as a source for derivation of multiple lineages of adult cells. In spite of these distinctive potentials, studies have raised concerns over the prolonged culture of ESCs. Formation of in vivo teratomas has been reported in implantation of in vitro cultured undifferentiated ESCs. Additionally, difficulties in finding patient-matched ESCs are an obstacle. Finally, because isolating ESCs involves the destruction or manipulation of pre-implantation stage embryo, there are lots of ethical controversies surrounding their usage[19].

iPSCs

iPSCs are generated via the induction of expression of certain genes in non-pluripotent adult cells. This technique was first developed in 2006 by Takahashi and Yamanaka, who introduced four transcription factors, Oct-4, c-Myc, KLF4, and SOX2, into mouse fibroblasts. These factors contributed to the maintenance of pluripotency in ESCs and are sufficient to generate ESC-like colonies[20]. A year later, Yamanaka improved on the reprogramming approach, leading to generation of iPSCs that were indistinguishable from ESCs[21]. Direct derivation of iPSCs from adult tissues not only helps to bypass the need for embryos as the pluripotent stem cell source but also makes personalized cell therapy a viable option. This method could generate unlimited supplies of young autologous pluripotent stem cells with a promising future in the field of regenerative medicine[22,23].

In contrast, several challenges still exist. Primarily, the efficacy of the reprogramming process is considerably low. For example, the rate at which somatic cells were reprogrammed into iPSCs in Yamanaka's original mouse study was 0.01%–0.1%. Although protocols for the induction of pluripotency are evolving, experimental evidence for appropriate initial cell type, transcription factor combinations, gene vectors, and methods of cell culture still lack the consistency required for clinical applications. In addition, induction of pluripotency and the process of reprogramming, itself causes genomic instability and adversely affects the cellular integrity. Moreover, the reprogramming factors (such as c-Myc) are known to be proto-oncogenes. Also, the retained epigenetic memory of the past somatic identity in newly generated iPSCs may influence the potency and in vivo functionality of engineered tissues[24]. However, recent rapid progress of several clinical studies have improved the outlook for this technology[25-27]. The first iPSC-derived therapy was done for age-related macular degeneration patients at Japan’s RIKEN Institute[28]. Encouraging results have smoothed the path for other scientific groups to seek clinical trials for the iPSC-based treatment of cardiac diseases, Parkinson’s disease, and blood clotting disorders[29-32].

Adult/somatic stem cells

Adult stem cells are populations of undifferentiated cells that unlike ESCs are found in mature tissues and organs throughout the postnatal life. These progenitor cells are responsible for tissue cell turnover and maintenance of injured tissues. Their easy accessibility, availability, and self-renewal ability introduce adult stem cells as a preferred cell source for transplantation. In spite of these great potentials, adult stem cells are not perfect. First, unlike pluripotent ESCs, adult stem cells are usually multipotent and can only give rise to a limited number of cell lineages of their specific tissue. Although adult stem cells can be obtained from both allogeneic and autologous sources, the age-dependent progressive deterioration of stem cell function is an important issue to be expected[11,33]. In addition, due to replicative senescence after prolonged culture periods, the proliferative ability of these cells declines rapidly[9,33].

Genetically-modified stem cells

Genetically modified stem cells are born out of the junction of two focus points of intense research: Gene therapy and stem cell therapy. Gene modification of cells prior to transplantation is one of the proposed solutions to overcome cell source challenges and to enhance cell proliferation and function[34]. Various gene therapy approaches are proposed, including the creation of genetically or epigenetically modified cells expressing useful proteins, growth factors or growth factor receptors, transcription factors, neurotransmitters and their receptors, and neuropeptides or creating cells that have the ability to recruit host cells to the implantation site[35,36].

The promise of using clustered regularly interspaced short palindromic repeats (CRISPR) technology brings about new hope as a tool for the gene editing of stem cells[37,38]. Brunger et al[39] used CRISPR for targeted deletion of the interleukin-1 receptor type 1 gene in murine iPSCs to make custom-made inflammation resistant cartilage cells. Genome editing by CRISPR has also been used to correct Duchenne muscular dystrophy patients derived iPSCs successfully to differentiate muscle cells that express functional protein[40]. Moreover, genetic modification of patient-derived iPSCs using CRISPR and other genetic engineering tools has been used for hemoglobinopathies such as β-thalassemia and sickle-cell anemia[41-44].

Immortalized cell lines are genetically mutated cells with unlimited propagation potential that are generated to prevail major challenges of cell source availability, such as early onset of cellular senescence and the consecutive limited cell expansion and differentiation potential. The mutations required for immortality can occur naturally or can be induced intentionally. There are several possible gene modification methods to bypass the senescence block. Viral oncogenes such as SV40 and E6/E7 proteins of oncogenic human papillomaviruses are used for regulating human cellular senescence[45-48]. One possible gene editing approach is artificial expression of key proteins required for immortality such as telomerase (discussed in later parts of this article). However, it was reported that they might be associated with genomic instability and increased risk of cell transformation[33,49,50].

Conditionally reprogrammed cells are another gene therapy approach used to rapidly and efficiently generate an unlimited number of patient-derived cells. In this technique by using both fibroblast feeder cells and a Rho-associated kinase inhibitor, Liu et al[51] indefinitely extend the life span of primary human keratinocytes ex vivo. Unlimited propagation of these karyotype-stable and non-tumorigenic cells offer opportunities for regenerative medicine as these cells have a stem cell-like phenotype[52].

In spite of preliminary success, several hurdles prevent both laboratory and particularly clinical applications of these gene-editing technologies. Genetic and epigenetic changes might cause unresolved issues to the patient. Transgenic genes, vector genes, or non-autologous stem cells might trigger immune reactions or even induce neoplastic transformation. In addition, developing an ideal gene vector system is next to impossible; the most common vectors are viruses. Beyond the uncontrollable insertional mutagenesis of viruses leading to increased risk of malignant transformation, viruses can cause adverse events such as toxicity and immune and inflammatory responses[34,53].

Ethical issues of gene editing of stem cells should not be overlooked. Matters like safety and efficacy of gene editing, including off-target mutations, raise concerns regarding human enhancement and eugenics that must be closely regulated. All in all, it is a necessity to set boundaries for techniques that have dire consequences[37,54].

CELLULAR AGING AS A LIMITING FACTOR

Aging, whether it is in a stem cell or a fully differentiated cell, seems to be the result of particularly shared processes. Some believe that aging occurs due to the incapability of senescent stem cells to contribute in tissue repair and regeneration, while others suggest that the vicious cycle of the dysfunctional relationship between stem cells and their niche cells is the leading contributor in the progressive deterioration during aging[55].

Cell intrinsic changes usually occur due to the accumulation of damage caused by normal cellular processes like metabolism and proliferation, while cell extrinsic changes are derived by a factor external to the cell subjected to those changes, such as paracrine and endocrine factors, ionizing radiation. and changes in the extracellular matrix[55]. Sometimes it is almost impossible to delineate the intrinsic and extrinsic changes. For instance, free radicals produced during both oxidative phosphorylation (a cell intrinsic factor) and generated by ionizing radiation (a cell extrinsic factor) can harm cellular components leading to senescence[56]. These intrinsic and extrinsic elements are discussed in detail in the following sections.

Functional decline of aged stem cells

During aging, several functional properties of the stem cells are being affected[55]. For instance, aged stem cells, especially neural stem cells[57-59], germline stem cells[60], and muscle satellite cells[61-65], lose their cellular polarity. As a consequence, they lose their ability to divide asymmetrically, a key feature of stem cells helping them to preserve the stem cell repertoires[55]. This loss of polarity is granted mostly by cell extrinsic factors like aged niche cells, dysfunctional adhesion molecules, disrupted morphogen, growth factors signaling, and inflammation[66-71]. Another phenomenon that is seen in aged stem cells is a lineage bias in the differentiation of their progenies. To enumerate, aged hematopoietic stem cells (HSCs) tend to skew toward the myeloid lineages more prominently compared to young and fully functional HSCs, a circumstance that leads to the incompetence of adaptive immune system in aged individuals[55,72-83].

Another example is loss of osteogenic differentiation and biased adipogenic commitment of mesenchymal stem cells (MSCs), which contributes to osteoporosis and bone marrow fat accumulation in aged individuals. Over-expression of receptor activator of nuclear factor kappa B ligand, down-regulation of peroxisome proliferator-activator receptor gamma, and suppression of forkhead box family O3 by protein kinase B (AKT) signaling in aged MSCs are proposed as the mechanisms responsible in this phenomenon[84-86]. This age-associated skewed differentiation is not completely understood for cells like ESCs and iPSCs. For instance, Xie et al[87] showed that H9 ESCs have an increased tendency for ectodermal lineages; however, this may be explained by the culture media composition. However, they observed no difference in teratoma formation between old and young ESCs. iPSCs were found to have a different story; while some studies have claimed that iPSCs have skewed differentiation capacity, probably because of their retained epigenetic memory of their original cell lines[88], other studies have reported that the iPSC differentiation capacity has no correlation with the cell source they are originated from[24,89].

Additionally, aged stem cells lose their migratory and homing potential, due to both cell intrinsic and cell extrinsic changes. For instance, transplantation of young mouse HSCs to old individuals delivers a lower yield compared to young recipients, which is due to the inferiority of the aged bone marrow niche[90,91]. Additionally, transplantation of old HSCs to young individuals is less effective in contrast to young donor cells, which shows a decline in the homing capacity of old stem cells due to intrinsic factors[72,90,91]. Another interesting aspect of HSCs is that using immunophenotyping, it has been shown that the number of HSCs increases with age; however, functional evaluation of these immunophenotypes shows reduced engraftment and improper differentiation in the new host[55,78-80,92]. This indicates a decrease in the population of functional HSCs or a form of clonality that happens with aging[55,81-83].

Yet, another example is heterochronic transplantation of aged mice muscle satellite cells into young recipients[61-65]. These cells show much lower regenerative capacity in old donors compared to young recipients, mostly due to cell extrinsic factors including, but not limited to, Wnt, Notch, and transforming growth factor beta signaling as well as altered Janus kinase-signal transducer and activator of transcription signal transduction[93,94]. In addition to the declined in vivo engraftment potential, aged satellite cells also have a reduced in vitro proliferation capacity[95-99].

Replicative senescence, Hayflick limit, and telomere length

Replicative senescence, equally known as Hayflick limit, is a phenomenon observed in vitro in which a primary cell or a stem cell stops dividing after a particular number of doublings. While the mechanism of Hayflick limit is not thoroughly understood, many attribute telomere attrition and genomic instability as the principal mechanism of replicative senescence[84,100]. For instance, MSCs stop dividing after 20 to 40 doublings when their telomeres are between 5.8 and 10.5 kb[84,101-103]. Despite the fact that telomere attrition can be considered an intrinsic change in stem cell replicative senescence, several questions remain to be addressed. Considering the fact that non-dividing cells also senesce[104-106], can we really take telomere attrition as one of the integral causes of aging, or is it just one of the many “effects” of the aging process that worsens this vicious cycle? Additionally, if telomere attrition is a deriving cause for aging, is it possible to increase the lifespan of a model organism, like a mouse or a rat, by “telomerization” or telomere lengthening? While the answers to these questions are controversial, it seems possible to immortalize cell lines via expression of telomerase subunits. Human telomerase reverse transcriptase (hTERT)-immortalized cells show extended life span with improved functional activities[58]. A successful example of this approach is the use of immortalized human keratinocyte cell lines in the treatment of chronic wounds and complex skin defects[59,60]. Notwithstanding, one study showed that over-expression of TERT only increases the median lifespan of the cancer-resistant mice, implying that telomere attrition might be important only in the late stages of aging[55,107,108]. Additionally, as mice have very long telomeres, it is not clear why they have a much shorter lifespan. Knocking out the RNA component of telomerase has no obvious life-threatening effect up to the sixth generation of these mice. Albeit, HSCs of the fourth generation started to show lineage skewness[109,110].

Stem cells spend most of their life in a quiescent state, probably to avoid the replicative damages, especially those related to DNA duplication. These quiescent stem cells are more likely to acquire destructive DNA damage after a double-strand break compared to a cell in its proliferative state. Quiescent stem cells mostly use the error-prone non-homologous end joining repair mechanism, while proliferative cells utilize homologous recombination, a much more accurate repair mechanism[111,112].

Autophagy

Perhaps one of the most extensively studied factors involved in aging is autophagy. Autophagy is a conserved mechanism that has evolved to recycle the damaged structures and organelles in a eukaryotic cell. This very sophisticated feature integrates the signal from several pathways to regulate the level of protein degradation. AMPK (adenosine monophosphate-dependent protein kinase), mTOR (mechanistic target of rapamycin), and ULK1 (Unc-51 like autophagy activating kinase 1) are the most important upstream signaling pathways of autophagy that regulate atg (autophagy related) genes and autophagosome formation. AMPK senses the ratio of AMP:ATP and activates ULK1 whenever the cell requires more energy. mTOR, on the other hand, inhibits ULK1 and autophagosome formation whenever it integrates the signal from nutrients and growth factors, which are the prerequisites of anabolism[11,113,114]. Every cell tries to strike a balance between the three forms of autophagy (macro-autophagy, micro-autophagy, and chaperone-mediated autophagy) and protein synthesis[115]. It has been shown that autophagy is decreased in aged cells of both animal models and humans, regardless of whether it is a stem cell or a fully differentiated one. While autophagy declines progressively, tiny amounts of damage gradually accumulate throughout time[115-117]. Stem cells have at least two mechanisms to prevent these damaged proteins and organelles to build-up: the asymmetric division (which diminishes with aging)[55,118-120] and maintaining high levels of autophagy and proteasome activity[55,121].

Aged MSCs and HSCs show accumulation of autophagic vesicles and inclusion bodies with LC3II or ubiquitin expression, which are the features of decreased autophagy with age. Rapamycin or spermidine treatment restores the autophagic capacity, leading to clearance of those accumulated autophagic vesicles and inclusions[115,122,123]. Additionally, Ho et al[123] showed that more than two-thirds of the HSCs in an aging population have very low levels of autophagy and skewed and escalated differentiation to myeloid lineages, while only less than a third of them had high levels of autophagy and regenerative potential comparable to that of young HSCs[115,123]. It has also been shown that while young MSCs in GFP-LC3 transgenic mice have high levels of autophagy in the quiescent state, this capacity fades with aging[115]. Furthermore, using conditional knock-out mice, when autophagosome formation is genetically compromised, senescence can rise from impairment of proteostasis[115,122]. These studies substantiate the importance of autophagy for maintaining stemness in a quiescent stem cell[115].

Epigenetic changes

In a fertilized egg, the genome of two mature and non-young individuals get stripped of epigenetic marks (except for the imprinted areas) to form a new very young individual with almost the same lifespan as the parents[124]. This epigenetic reprogramming can be emulated by in vitro induction of totipotency/pluripotency. The nucleus of a somatic cell can either be fused with the cytoplasm of an enucleated oocyte (somatic cell nuclear transfer) or be transfected with viruses expressing Yamanaka factors (SOX2, c-Myc, Oct-4, and KLF4)[20] to produce an iPSC[55,125-128]. Although this reprogrammed cell is very similar to a freshly young ESC in many aspects, its epigenome is slightly different[127]. In fact, iPSCs are reprogrammed with regard to the age-, tissue-, and senescence-associated DNA methylation patterns but keep some donor-specific DNA methylation patterns[129]. In 2013, Abad et al[130] produced a transgenic mouse expressing the four Yamanaka factors in every cell upon administration of doxycycline. These mice usually develop teratomas in several organs and tissues. Another interesting fact about epigenetic reprograming is that it is possible to make phenotypically young neural stem cells from iPSCs, which are generated from aged fibroblast, while direct transdifferentiating neural stem cells from aged fibroblasts maintains the aged phenotype[55,131].

This reprogramability of the epigenome helps us to unravel detailed mechanisms of aging in order to find a way to conquer it in the future; however, it is obviously not a practical formula for rejuvenation[55]. Perhaps it is better to use epigenomic results to reinforce gene regulatory networks and to decipher what signals are differentially active in old cells compared to their younger counterparts.

miRNA

Among the differentially expressed genes in aged cells in comparison with the young cells, there are several non-coding RNAs, including miRNAs expressed (some of which are even proposed as biomarkers for aging)[56,84,132]. miR-195 is over-expressed in aged cells and reduces the telomerase reverse transcriptase; also, knocking-down this miRNA in MSCs increases their regenerative capacity when transplanted to the infarcted myocardium[56,133]. Another example is over-expressed miR-34a, which is elevated in infarcted mouse hearts and is associated with apoptosis and senescence. Also, its inhibition decreases the number of apoptotic cells in cardiac tissue[56,134,135]. Some of these differentially expressed miRNAs control proliferative and regenerative capacity of the stem cells by regulating cell cycle transition and stemness factors (such as Nanog)[84,133,136-138].

Role of mitochondria in stem cell senescence

Free radicals, otherwise known as reactive oxygen species (ROS), are a well-recognized origin of age-related molecular injuries including but not limited to nuclear and mitochondrial DNA mutations, organelle damages, and lipofuscins. Chronic inflammation, ionizing radiation, and mitochondrial dysfunction are the most prominent sources of ROS in cells[56,139,140]. Stem cells employ several mechanisms to keep ROS and its damage at bay. To enumerate, quiescent HSCs depend predominantly on glycolysis to limit ROS production[75,141].

Sirtuins, (SIRT1–SIRT7) a conserved family of NAD+-dependent deacetylases of which SIRT1 is the best known, appears to increase mitochondrial turnover by activation of mitophagy. Activation of SIRTs can considerably extend the replicative capacity of human bone marrow stem cells[142] and human fibroblasts[11,143]. Additionally, SIRTs boost the stress-relieving and antioxidant mechanisms in cells. Studies show that over-expression of certain SIRTs increases catalase and superoxide dismutase, while their knock-down compromises cell proliferation and increases cellular senescence[56,144-147].

On the other hand, stem cells have a unique way to get rid of the damaged proteins and organelles[55,75]. During asymmetrical division, stem cells actively accumulate these injuries in the differentiating daughter cell, while keeping the daughter stem cell almost clean of them[109,118-120,148]. This polarized division is lost in certain stem cells, like HSCs and germline stem cells, during aging. As a matter of fact, less-polarized HSCs are more biased toward myeloid lineages[55,70,71]. Furthermore, autophagy and proteasome-mediated degradation, the other mechanisms that keep even the quiescent stem cells clean, diminish with aging. When dysfunctional mitochondria cannot be recycled by mitophagy (macro-autophagy of mitochondria), it generates more ROS. As a result, we observe a vicious cycle between impaired autophagy, mitochondrial dysfunction, and ROS-mediated injuries[56,139,140,149]. Xie et al[87] found that the most prominent changes that occur in long-term passaged ESCs have to do with the mitochondria; older passages of H9 and PKU1 hESC lines have elevated mitochondrial mass, ROS level, and mitochondrial membrane potential. On the other hand, aged iPSCs develop defects in their nuclear envelop[150], which might be the cause of interference in SIRT and NF-kB nuclear transportation and downstream signaling in these cells[151,152].

Above all, some studies contradict ROS as a contributing factor in aging. At the cellular level, Zhu et al[153] showed that there is “no evident dose-response effect between cellular ROS level and its cytotoxicity.” For instance, they showed that while all three of the piperlongumin, beta-phenylethyl isothiocyanate, and lactic acid increased ROS in the cultured cells, only piperlongumin and beta-phenylethyl isothiocyanate, two ROS-based chemotherapeutic agents, killed the cells and lactic acid “spared them.” Additionally, although chemical depletion of glutathione increased ROS much higher than piperlongumin and beta-phenylethyl isothiocyanate, it did not affect the cell growth in cultured samples. However, these results were achieved in cancer cells, and it is unclear if similar mechanisms also happen in stem cells. Le Gal et al[154] showed that administration of the antioxidant N-acetyl cysteine to a mouse model of melanoma not only decreased the survival of the mice but also increased the severity of their tumors by increasing metastasis. Biesalski et al[155] meta-analytically reevaluated clinical effectiveness of antioxidants on mortality and health. They showed that micronutrients, including those with antioxidant activity, are only effective in those with the deficiencies or the risk of deficiencies, but not effective in individuals with the micronutrients above the minimum required level. All in all, these counterexamples provide sufficient evidence to raise a reasonable doubt toward ROS-based therapeutics.

INTERVENTIONS FOR REJUVENATION OF AGED STEM CELLS

The ultimate goal of unraveling mechanisms of aging is geroprotection (preventing from aging) or rejuvenation (making a senescent cell young again). For this purpose, there are three options: Changing the extrinsic factors, altering the intrinsic factors, or manipulating the genomic targets of those changes. We can either use pharmacological means, modify the environment in which the stem cells reside, or genetically manipulate the stem cells.

Using pharmacological means to prevent aging or even rejuvenate is, perhaps, the most practical measure. Different mechanisms have been targeted pharmacologically. For instance, antioxidants like vitamin C and N-acetyl cysteine have been used to reduce ROS both in vitro and in vivo[75,154,156]. However, their efficacy is limited, especially in vivo. Although antioxidants to some extent show geroprotection in cell culture, the possible life extension by reduction of free radicals is challenged by in vivo experiments[115,153-155]. Comparatively, SIRTs are another example of drug targets for geroprotection. Resveratrol, resveratrol-mimicking compounds, and NAD+ seem to hinder aging both in vitro and in vivo by activating certain members of the SIRT family[157-160]. In particular, resveratrol improves metabolism and enhances DNA repair, which are critically important in aging[11,157].

Metabolic dysfunction is yet another focus for research on geroprotection. Rapamycin, spermidine, quercetin, and metformin are only a few examples of the drugs that increase the lifespan by this mechanism, whether it is in cell culture or in vivo[56,161]. As we previously discussed, the balance between protein synthesis and protein recycling is disrupted in aging. Rapamycin that inhibits mTOR gives an advantage to the autophagic side of the balance between autophagy and protein synthesis[56,114,162]. Likewise, metformin activates AMPK to increase autophagy and other anti-stress mechanisms in the cell and slightly inhibits mTOR complex 1 through targets upstream of mTOR complex 1[11,163,164]. This boosted autophagy helps the cell to get rid of the damaged organelles and macromolecules much faster than it did before. Thus, it delays the damage accumulation, which is an important factor in dysfunctioning of stem cells[73,115,165-167]. Both of these small molecules have been shown to increase the lifespan, decrease the doubling time, and improve functional properties of stem cells, e.g., engraftment, migratory, and regenerative potential[11,115,165-167]. Furthermore, combination treatment of cardiac stem cells with rapamycin and resveratrol improves the cardiac output of the infarcted myocardium in mice[56,168].

Caloric restriction (CR) is the most effective intervention for lifespan extension. Mechanistically, CR exerts its benefits through the alteration of the nutrient/growth factor-sensing mTOR signaling, energy-sensing AMPK signaling, stress-fighting forkhead box family O signaling, and SIRTs[55,169-172]. Therefore, CR not only increases the longevity of stem cells but also enhances the performance of niche cells that support stem cells[55,173-176]. Interestingly, most of the compounds that extend the lifespan by improving metabolism, especially those that promote autophagy like metformin and spermidine, are known as CR mimetics[73,174,177,178]. In fact, the geroprotective activity of CR is repressed by hindering autophagy[73,117]. While CR works best in vivo, it is not a practical way to extend the in vitro lifespan of stem cells, precisely because it limits the doubling time of cells. Thus, these CR mimetics might be the most practical intervention to be used in cell culture[11].

Although interventions like genetic manipulation might effectively work to counteract senescence in stem cells, their cost and safety concerns limit their application[75,179]. Studies mentioned interventions like over-expression of telomerase as a proposed mechanism for counteracting replicative senescence in MSCs[84,180]. For instance, over-expression of hTERT in MSCs increased their lifespan, while the normal karyotype was maintained[84,180,181]. Another strategy to genetically prevent aging is knocking-down either the retinoblastoma protein gene or the p16INK4a gene[84,182,183]. Retinoblastoma gene silencing decreases the age-related DNA damage and senescence as well as increases the functionality of MSCs[84,182]. Finally, manipulating miRNAs could be an effective strategy, but it needs further experimental support. To enumerate, knock-down of miR-195 leads to increased expression of hTERT, and forkhead box family O3 also intensified phosphorylation of protein kinase B (AKT) in senescent MSCs[84,133].

CONCLUSION

The way toward the production of tissue engineered products still has serious hurdles to overcome: The choice of cell source, proper biomaterial selection, maintaining blood supply by designing suitable scaffolds, and three-dimensional tissue architecture. Combined efforts to prevail over these major obstacles are warranted to pave the way for achieving tissue engineered products at a commercial scale.

With regards to the choice of cell source, aging is a limiting factor. Aging, as inevitable as it seems, is proven to be conquerable. In different cell types the problem of aging is preventable and to some extent reversible. As aging is a very complex and dynamic phenomenon, it would be better to approach it from a systems biology point of view to reach the best results. Perhaps we need to target multiple pathways to find the maximum efficacy. Regardless of the application of the stem cells, i.e. tissue engineering and cell therapy, we have to overcome aging, both in the original cell source and in the in vitro proliferation.

ACKNOWLEDGEMENTS

The authors acknowledge Dr. Seyed Mojtaba Hosseini and Dr. Sina Kardeh for their insights and comments on the manuscript. The authors wish to thank Mr. H. Argasi at the Research Consultation Center (RCC) of Shiraz University of Medical Sciences for his invaluable assistance in editing this manuscript.

Footnotes

Manuscript source: Invited manuscript

Specialty type: Cell and tissue engineering

Country of origin: Iran

Peer-review report classification

Grade A (Excellent): 0

Grade B (Very good): B

Grade C (Good): C, C

Grade D (Fair): D

Grade E (Poor): 0

P-Reviewer: Alonso MBD, Binetruy B, Bragança J, Sonntag KC S-Editor: Ma YJ L-Editor: Filipodia E-Editor: Qi LL

References
1.  Vacanti JP, Vacanti CA.  The history and scope of tissue engineering. In: Principles of Tissue Engineering (Fourth Edition). Elsevier Inc 2014; 3-8.  [PubMed]  [DOI]  [Cited in This Article: ]
2.  Khademhosseini A, Vacanti JP, Langer R. Progress in tissue engineering. Sci Am. 2009;300:64-71.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 299]  [Cited by in F6Publishing: 250]  [Article Influence: 16.7]  [Reference Citation Analysis (0)]
3.  Langer R, Vacanti JP. Tissue engineering. Science. 1993;260:920-926.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 8352]  [Cited by in F6Publishing: 6440]  [Article Influence: 207.7]  [Reference Citation Analysis (0)]
4.  Hoffman T, Khademhosseini A, Langer R. Chasing the Paradigm: Clinical Translation of 25 Years of Tissue Engineering. Tissue Eng Part A. 2019;25:679-687.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 74]  [Cited by in F6Publishing: 64]  [Article Influence: 12.8]  [Reference Citation Analysis (0)]
5.  Langer R. Perspectives and challenges in tissue engineering and regenerative medicine. Adv Mater. 2009;21:3235-3236.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 126]  [Cited by in F6Publishing: 100]  [Article Influence: 6.7]  [Reference Citation Analysis (0)]
6.  Bakhshandeh B, Zarrintaj P, Oftadeh MO, Keramati F, Fouladiha H, Sohrabi-Jahromi S, Ziraksaz Z. Tissue engineering; strategies, tissues, and biomaterials. Biotechnol Genet Eng Rev. 2017;33:144-172.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 84]  [Cited by in F6Publishing: 86]  [Article Influence: 14.3]  [Reference Citation Analysis (0)]
7.  Jara Avaca M, Gruh I. Bioengineered Cardiac Tissue Based on Human Stem Cells for Clinical Application. Adv Biochem Eng Biotechnol. 2018;163:117-146.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1]  [Cited by in F6Publishing: 1]  [Article Influence: 0.2]  [Reference Citation Analysis (0)]
8.  Vunjak-Novakovic G, Tandon N, Godier A, Maidhof R, Marsano A, Martens TP, Radisic M. Challenges in cardiac tissue engineering. Tissue Eng Part B Rev. 2010;16:169-187.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 365]  [Cited by in F6Publishing: 320]  [Article Influence: 22.9]  [Reference Citation Analysis (0)]
9.  Murtuza B, Nichol JW, Khademhosseini A. Micro- and nanoscale control of the cardiac stem cell niche for tissue fabrication. Tissue Eng Part B Rev. 2009;15:443-454.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 67]  [Cited by in F6Publishing: 69]  [Article Influence: 4.9]  [Reference Citation Analysis (0)]
10.  Dixit P, Katare R. Challenges in identifying the best source of stem cells for cardiac regeneration therapy. Stem Cell Res Ther. 2015;6:26.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 75]  [Cited by in F6Publishing: 80]  [Article Influence: 8.9]  [Reference Citation Analysis (0)]
11.  Khorraminejad-Shirazi M, Farahmandnia M, Kardeh B, Estedlal A, Kardeh S, Monabati A. Aging and stem cell therapy: AMPK as an applicable pharmacological target for rejuvenation of aged stem cells and achieving higher efficacy in stem cell therapy. Hematol Oncol Stem Cell Ther. 2018;11:189-194.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17]  [Cited by in F6Publishing: 15]  [Article Influence: 2.1]  [Reference Citation Analysis (0)]
12.  Sanz-Ruiz R, Fernández-Avilés F. Autologous and allogeneic cardiac stem cell therapy for cardiovascular diseases. Pharmacol Res. 2018;127:92-100.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 19]  [Cited by in F6Publishing: 19]  [Article Influence: 2.7]  [Reference Citation Analysis (0)]
13.  Efimenko AY, Kochegura TN, Akopyan ZA, Parfyonova YV. Autologous Stem Cell Therapy: How Aging and Chronic Diseases Affect Stem and Progenitor Cells. Biores Open Access. 2015;4:26-38.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 51]  [Cited by in F6Publishing: 57]  [Article Influence: 6.3]  [Reference Citation Analysis (0)]
14.  Langer RS, Vacanti JP. Tissue engineering: the challenges ahead. Sci Am. 1999;280:86-89.  [PubMed]  [DOI]  [Cited in This Article: ]
15.  Singh AK, McGuirk JP. Allogeneic Stem Cell Transplantation: A Historical and Scientific Overview. Cancer Res. 2016;76:6445-6451.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 111]  [Cited by in F6Publishing: 121]  [Article Influence: 15.1]  [Reference Citation Analysis (0)]
16.  Cooper DK. A brief history of cross-species organ transplantation. Proc (Bayl Univ Med Cent). 2012;25:49-57.  [PubMed]  [DOI]  [Cited in This Article: ]
17.  Baharvand H, Ashtiani SK, Taee A, Massumi M, Valojerdi MR, Yazdi PE, Moradi SZ, Farrokhi A. Generation of new human embryonic stem cell lines with diploid and triploid karyotypes. Dev Growth Differ. 2006;48:117-128.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 117]  [Cited by in F6Publishing: 121]  [Article Influence: 8.1]  [Reference Citation Analysis (0)]
18.  Sierra RA, Hoverter NP, Ramirez RN, Vuong LM, Mortazavi A, Merrill BJ, Waterman ML, Donovan PJ. TCF7L1 suppresses primitive streak gene expression to support human embryonic stem cell pluripotency. Development. 2018;145:dev161075.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 11]  [Cited by in F6Publishing: 13]  [Article Influence: 2.2]  [Reference Citation Analysis (0)]
19.  Macrin D, Joseph JP, Pillai AA, Devi A. Eminent Sources of Adult Mesenchymal Stem Cells and Their Therapeutic Imminence. Stem Cell Rev. 2017;13:741-756.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 54]  [Cited by in F6Publishing: 59]  [Article Influence: 9.8]  [Reference Citation Analysis (0)]
20.  Takahashi K, Yamanaka S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell. 2006;126:663-676.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17989]  [Cited by in F6Publishing: 17055]  [Article Influence: 947.5]  [Reference Citation Analysis (0)]
21.  Takahashi K, Tanabe K, Ohnuki M, Narita M, Ichisaka T, Tomoda K, Yamanaka S. Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell. 2007;131:861-872.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 14327]  [Cited by in F6Publishing: 13579]  [Article Influence: 848.7]  [Reference Citation Analysis (0)]
22.  Shi Y, Inoue H, Wu JC, Yamanaka S. Induced pluripotent stem cell technology: a decade of progress. Nat Rev Drug Discov. 2017;16:115-130.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 960]  [Cited by in F6Publishing: 851]  [Article Influence: 121.6]  [Reference Citation Analysis (0)]
23.  de Lázaro I, Yilmazer A, Kostarelos K. Induced pluripotent stem (iPS) cells: a new source for cell-based therapeutics? J Control Release. 2014;185:37-44.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 48]  [Cited by in F6Publishing: 53]  [Article Influence: 5.3]  [Reference Citation Analysis (0)]
24.  Youssef AA, Ross EG, Bolli R, Pepine CJ, Leeper NJ, Yang PC. The Promise and Challenge of Induced Pluripotent Stem Cells for Cardiovascular Applications. JACC Basic Transl Sci. 2016;1:510-523.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 32]  [Cited by in F6Publishing: 34]  [Article Influence: 4.3]  [Reference Citation Analysis (0)]
25.  Kimbrel EA, Lanza R. Current status of pluripotent stem cells: moving the first therapies to the clinic. Nat Rev Drug Discov. 2015;14:681-692.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 191]  [Cited by in F6Publishing: 170]  [Article Influence: 18.9]  [Reference Citation Analysis (0)]
26.  Velmurugan BK, Bharathi Priya L, Poornima P, Lee LJ, Baskaran R. Biomaterial aided differentiation and maturation of induced pluripotent stem cells. J Cell Physiol. 2019;234:8443-8454.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 6]  [Cited by in F6Publishing: 7]  [Article Influence: 1.2]  [Reference Citation Analysis (0)]
27.  Kwon SG, Kwon YW, Lee TW, Park GT, Kim JH. Recent advances in stem cell therapeutics and tissue engineering strategies. Biomater Res. 2018;22:36.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 80]  [Cited by in F6Publishing: 85]  [Article Influence: 14.2]  [Reference Citation Analysis (0)]
28.  Mandai M, Watanabe A, Kurimoto Y, Hirami Y, Morinaga C, Daimon T, Fujihara M, Akimaru H, Sakai N, Shibata Y, Terada M, Nomiya Y, Tanishima S, Nakamura M, Kamao H, Sugita S, Onishi A, Ito T, Fujita K, Kawamata S, Go MJ, Shinohara C, Hata KI, Sawada M, Yamamoto M, Ohta S, Ohara Y, Yoshida K, Kuwahara J, Kitano Y, Amano N, Umekage M, Kitaoka F, Tanaka A, Okada C, Takasu N, Ogawa S, Yamanaka S, Takahashi M. Autologous Induced Stem-Cell-Derived Retinal Cells for Macular Degeneration. N Engl J Med. 2017;376:1038-1046.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 910]  [Cited by in F6Publishing: 856]  [Article Influence: 122.3]  [Reference Citation Analysis (0)]
29.  Li W, Chen S, Li JY. Human induced pluripotent stem cells in Parkinson's disease: A novel cell source of cell therapy and disease modeling. Prog Neurobiol. 2015;134:161-177.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 20]  [Cited by in F6Publishing: 23]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
30.  Xiao B, Ng HH, Takahashi R, Tan EK. Induced pluripotent stem cells in Parkinson's disease: scientific and clinical challenges. J Neurol Neurosurg Psychiatry. 2016;87:697-702.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 37]  [Cited by in F6Publishing: 35]  [Article Influence: 4.4]  [Reference Citation Analysis (0)]
31.  Feng Q, Shabrani N, Thon JN, Huo H, Thiel A, Machlus KR, Kim K, Brooks J, Li F, Luo C, Kimbrel EA, Wang J, Kim KS, Italiano J, Cho J, Lu SJ, Lanza R. Scalable generation of universal platelets from human induced pluripotent stem cells. Stem Cell Reports. 2014;3:817-831.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 166]  [Cited by in F6Publishing: 167]  [Article Influence: 16.7]  [Reference Citation Analysis (0)]
32.  Yoshida Y, Yamanaka S. Induced Pluripotent Stem Cells 10 Years Later: For Cardiac Applications. Circ Res. 2017;120:1958-1968.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 170]  [Cited by in F6Publishing: 177]  [Article Influence: 25.3]  [Reference Citation Analysis (0)]
33.  Behrens A, van Deursen JM, Rudolph KL, Schumacher B. Impact of genomic damage and ageing on stem cell function. Nat Cell Biol. 2014;16:201-207.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 161]  [Cited by in F6Publishing: 152]  [Article Influence: 15.2]  [Reference Citation Analysis (0)]
34.  Phillips MI, Tang YL. Genetic modification of stem cells for transplantation. Adv Drug Deliv Rev. 2008;60:160-172.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 52]  [Cited by in F6Publishing: 53]  [Article Influence: 3.3]  [Reference Citation Analysis (0)]
35.  Yu H, Lu K, Zhu J, Wang J. Stem cell therapy for ischemic heart diseases. Br Med Bull. 2017;121:135-154.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 83]  [Cited by in F6Publishing: 94]  [Article Influence: 13.4]  [Reference Citation Analysis (0)]
36.  Dixit S, Baganizi DR, Sahu R, Dosunmu E, Chaudhari A, Vig K, Pillai SR, Singh SR, Dennis VA. Immunological challenges associated with artificial skin grafts: available solutions and stem cells in future design of synthetic skin. J Biol Eng. 2017;11:49.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 42]  [Cited by in F6Publishing: 44]  [Article Influence: 6.3]  [Reference Citation Analysis (0)]
37.  Kim EJ, Kang KH, Ju JH. CRISPR-Cas9: a promising tool for gene editing on induced pluripotent stem cells. Korean J Intern Med. 2017;32:42-61.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 39]  [Cited by in F6Publishing: 38]  [Article Influence: 5.4]  [Reference Citation Analysis (0)]
38.  Pulgarin DA, Nyberg WA, Bowlin GL, Espinosa A. CRISPR/Cas Systems in Tissue Engineering: A Succinct Overview of Current Use and Future Opportunities. Curr Trends Biomedical Eng Biosci. 2017;5:555670.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3]  [Cited by in F6Publishing: 2]  [Article Influence: 0.3]  [Reference Citation Analysis (0)]
39.  Brunger JM, Zutshi A, Willard VP, Gersbach CA, Guilak F. CRISPR/Cas9 Editing of Murine Induced Pluripotent Stem Cells for Engineering Inflammation-Resistant Tissues. Arthritis Rheumatol. 2017;69:1111-1121.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 45]  [Cited by in F6Publishing: 46]  [Article Influence: 6.6]  [Reference Citation Analysis (0)]
40.  Li HL, Fujimoto N, Sasakawa N, Shirai S, Ohkame T, Sakuma T, Tanaka M, Amano N, Watanabe A, Sakurai H, Yamamoto T, Yamanaka S, Hotta A. Precise correction of the dystrophin gene in duchenne muscular dystrophy patient induced pluripotent stem cells by TALEN and CRISPR-Cas9. Stem Cell Reports. 2015;4:143-154.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 381]  [Cited by in F6Publishing: 371]  [Article Influence: 37.1]  [Reference Citation Analysis (0)]
41.  Sebastiano V, Maeder ML, Angstman JF, Haddad B, Khayter C, Yeo DT, Goodwin MJ, Hawkins JS, Ramirez CL, Batista LF, Artandi SE, Wernig M, Joung JK. In situ genetic correction of the sickle cell anemia mutation in human induced pluripotent stem cells using engineered zinc finger nucleases. Stem Cells. 2011;29:1717-1726.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 252]  [Cited by in F6Publishing: 264]  [Article Influence: 22.0]  [Reference Citation Analysis (0)]
42.  Sun N, Zhao H. Seamless correction of the sickle cell disease mutation of the HBB gene in human induced pluripotent stem cells using TALENs. Biotechnol Bioeng. 2014;111:1048-1053.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 98]  [Cited by in F6Publishing: 103]  [Article Influence: 9.4]  [Reference Citation Analysis (0)]
43.  Huang X, Wang Y, Yan W, Smith C, Ye Z, Wang J, Gao Y, Mendelsohn L, Cheng L. Production of Gene-Corrected Adult Beta Globin Protein in Human Erythrocytes Differentiated from Patient iPSCs After Genome Editing of the Sickle Point Mutation. Stem Cells. 2015;33:1470-1479.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 139]  [Cited by in F6Publishing: 140]  [Article Influence: 17.5]  [Reference Citation Analysis (0)]
44.  Xie F, Ye L, Chang JC, Beyer AI, Wang J, Muench MO, Kan YW. Seamless gene correction of β-thalassemia mutations in patient-specific iPSCs using CRISPR/Cas9 and piggyBac. Genome Res. 2014;24:1526-1533.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 300]  [Cited by in F6Publishing: 296]  [Article Influence: 29.6]  [Reference Citation Analysis (0)]
45.  Shay J, Wright W. Both Rb and P53 are potential regulators of human cellular senescence. Hereditary Tumors. 1991;83:173-82.  [PubMed]  [DOI]  [Cited in This Article: ]
46.  Van der Haegen BA, Shay JW. Immortalization of human mammary epithelial cells by SV40 large T-antigen involves a two step mechanism. In Vitro Cell Dev Biol. 1993;29A:180-182.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 20]  [Cited by in F6Publishing: 22]  [Article Influence: 0.7]  [Reference Citation Analysis (0)]
47.  Liu X, Dakic A, Chen R, Disbrow GL, Zhang Y, Dai Y, Schlegel R. Cell-restricted immortalization by human papillomavirus correlates with telomerase activation and engagement of the hTERT promoter by Myc. J Virol. 2008;82:11568-11576.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 46]  [Cited by in F6Publishing: 48]  [Article Influence: 3.0]  [Reference Citation Analysis (0)]
48.  Hawley-Nelson P, Vousden KH, Hubbert NL, Lowy DR, Schiller JT. HPV16 E6 and E7 proteins cooperate to immortalize human foreskin keratinocytes. EMBO J. 1989;8:3905-3910.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 661]  [Cited by in F6Publishing: 605]  [Article Influence: 17.8]  [Reference Citation Analysis (0)]
49.  Centanni JM, Straseski JA, Wicks A, Hank JA, Rasmussen CA, Lokuta MA, Schurr MJ, Foster KN, Faucher LD, Caruso DM, Comer AR, Allen-Hoffmann BL. StrataGraft skin substitute is well-tolerated and is not acutely immunogenic in patients with traumatic wounds: results from a prospective, randomized, controlled dose escalation trial. Ann Surg. 2011;253:672-683.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 64]  [Cited by in F6Publishing: 66]  [Article Influence: 5.1]  [Reference Citation Analysis (0)]
50.  Kassem M, Abdallah BM, Yu Z, Ditzel N, Burns JS. The use of hTERT-immortalized cells in tissue engineering. Cytotechnology. 2004;45:39-46.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 25]  [Cited by in F6Publishing: 20]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
51.  Liu X, Ory V, Chapman S, Yuan H, Albanese C, Kallakury B, Timofeeva OA, Nealon C, Dakic A, Simic V, Haddad BR, Rhim JS, Dritschilo A, Riegel A, McBride A, Schlegel R. ROCK inhibitor and feeder cells induce the conditional reprogramming of epithelial cells. Am J Pathol. 2012;180:599-607.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 539]  [Cited by in F6Publishing: 562]  [Article Influence: 43.2]  [Reference Citation Analysis (0)]
52.  Suprynowicz FA, Upadhyay G, Krawczyk E, Kramer SC, Hebert JD, Liu X, Yuan H, Cheluvaraju C, Clapp PW, Boucher RC, Kamonjoh CM, Randell SH, Schlegel R. Conditionally reprogrammed cells represent a stem-like state of adult epithelial cells. Proc Natl Acad Sci U S A. 2012;109:20035-20040.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 203]  [Cited by in F6Publishing: 213]  [Article Influence: 17.8]  [Reference Citation Analysis (0)]
53.  Madry H, Cucchiarini M. Clinical potential and challenges of using genetically modified cells for articular cartilage repair. Croat Med J. 2011;52:245-261.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 30]  [Cited by in F6Publishing: 31]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
54.  Lau RW, Wang B, Ricardo SD. Gene editing of stem cells for kidney disease modelling and therapeutic intervention. Nephrology (Carlton). 2018;23:981-990.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 6]  [Cited by in F6Publishing: 7]  [Article Influence: 1.4]  [Reference Citation Analysis (0)]
55.  Schultz MB, Sinclair DA. When stem cells grow old: phenotypes and mechanisms of stem cell aging. Development. 2016;143:3-14.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 199]  [Cited by in F6Publishing: 195]  [Article Influence: 24.4]  [Reference Citation Analysis (0)]
56.  Kaur G, Cai C. Current Progress in the Rejuvenation of Aging Stem/Progenitor Cells for Improving the Therapeutic Effectiveness of Myocardial Repair. Stem Cells Int. 2018;2018:9308301.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 12]  [Cited by in F6Publishing: 12]  [Article Influence: 2.0]  [Reference Citation Analysis (0)]
57.  Katsimpardi L, Litterman NK, Schein PA, Miller CM, Loffredo FS, Wojtkiewicz GR, Chen JW, Lee RT, Wagers AJ, Rubin LL. Vascular and neurogenic rejuvenation of the aging mouse brain by young systemic factors. Science. 2014;344:630-634.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 684]  [Cited by in F6Publishing: 725]  [Article Influence: 72.5]  [Reference Citation Analysis (0)]
58.  Villeda SA, Luo J, Mosher KI, Zou B, Britschgi M, Bieri G, Stan TM, Fainberg N, Ding Z, Eggel A, Lucin KM, Czirr E, Park JS, Couillard-Després S, Aigner L, Li G, Peskind ER, Kaye JA, Quinn JF, Galasko DR, Xie XS, Rando TA, Wyss-Coray T. The ageing systemic milieu negatively regulates neurogenesis and cognitive function. Nature. 2011;477:90-94.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1110]  [Cited by in F6Publishing: 1234]  [Article Influence: 94.9]  [Reference Citation Analysis (0)]
59.  Villeda SA, Plambeck KE, Middeldorp J, Castellano JM, Mosher KI, Luo J, Smith LK, Bieri G, Lin K, Berdnik D, Wabl R, Udeochu J, Wheatley EG, Zou B, Simmons DA, Xie XS, Longo FM, Wyss-Coray T. Young blood reverses age-related impairments in cognitive function and synaptic plasticity in mice. Nat Med. 2014;20:659-663.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 650]  [Cited by in F6Publishing: 711]  [Article Influence: 71.1]  [Reference Citation Analysis (0)]
60.  Ryu BY, Orwig KE, Oatley JM, Avarbock MR, Brinster RL. Effects of aging and niche microenvironment on spermatogonial stem cell self-renewal. Stem Cells. 2006;24:1505-1511.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 193]  [Cited by in F6Publishing: 188]  [Article Influence: 10.4]  [Reference Citation Analysis (0)]
61.  Brack AS, Conboy MJ, Roy S, Lee M, Kuo CJ, Keller C, Rando TA. Increased Wnt signaling during aging alters muscle stem cell fate and increases fibrosis. Science. 2007;317:807-810.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1084]  [Cited by in F6Publishing: 1083]  [Article Influence: 63.7]  [Reference Citation Analysis (0)]
62.  Carlson BM, Faulkner JA. Muscle transplantation between young and old rats: age of host determines recovery. Am J Physiol. 1989;256:C1262-C1266.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 302]  [Cited by in F6Publishing: 277]  [Article Influence: 7.9]  [Reference Citation Analysis (0)]
63.  Chakkalakal JV, Jones KM, Basson MA, Brack AS. The aged niche disrupts muscle stem cell quiescence. Nature. 2012;490:355-360.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 555]  [Cited by in F6Publishing: 574]  [Article Influence: 47.8]  [Reference Citation Analysis (0)]
64.  Conboy IM, Conboy MJ, Wagers AJ, Girma ER, Weissman IL, Rando TA. Rejuvenation of aged progenitor cells by exposure to a young systemic environment. Nature. 2005;433:760-764.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1544]  [Cited by in F6Publishing: 1534]  [Article Influence: 80.7]  [Reference Citation Analysis (0)]
65.  Sinha M, Jang YC, Oh J, Khong D, Wu EY, Manohar R, Miller C, Regalado SG, Loffredo FS, Pancoast JR, Hirshman MF, Lebowitz J, Shadrach JL, Cerletti M, Kim MJ, Serwold T, Goodyear LJ, Rosner B, Lee RT, Wagers AJ. Restoring systemic GDF11 levels reverses age-related dysfunction in mouse skeletal muscle. Science. 2014;344:649-652.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 575]  [Cited by in F6Publishing: 605]  [Article Influence: 60.5]  [Reference Citation Analysis (0)]
66.  Pan L, Chen S, Weng C, Call G, Zhu D, Tang H, Zhang N, Xie T. Stem cell aging is controlled both intrinsically and extrinsically in the Drosophila ovary. Cell Stem Cell. 2007;1:458-469.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 162]  [Cited by in F6Publishing: 157]  [Article Influence: 9.8]  [Reference Citation Analysis (0)]
67.  Toledano H, D'Alterio C, Czech B, Levine E, Jones DL. The let-7-Imp axis regulates ageing of the Drosophila testis stem-cell niche. Nature. 2012;485:605-610.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 138]  [Cited by in F6Publishing: 131]  [Article Influence: 10.9]  [Reference Citation Analysis (0)]
68.  Wallenfang MR, Nayak R, DiNardo S. Dynamics of the male germline stem cell population during aging of Drosophila melanogaster. Aging Cell. 2006;5:297-304.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 102]  [Cited by in F6Publishing: 107]  [Article Influence: 5.9]  [Reference Citation Analysis (0)]
69.  Boyle M, Wong C, Rocha M, Jones DL. Decline in self-renewal factors contributes to aging of the stem cell niche in the Drosophila testis. Cell Stem Cell. 2007;1:470-478.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 206]  [Cited by in F6Publishing: 203]  [Article Influence: 12.7]  [Reference Citation Analysis (0)]
70.  Cheng J, Türkel N, Hemati N, Fuller MT, Hunt AJ, Yamashita YM. Centrosome misorientation reduces stem cell division during ageing. Nature. 2008;456:599-604.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 269]  [Cited by in F6Publishing: 277]  [Article Influence: 17.3]  [Reference Citation Analysis (0)]
71.  Köhler A, Schmithorst V, Filippi MD, Ryan MA, Daria D, Gunzer M, Geiger H. Altered cellular dynamics and endosteal location of aged early hematopoietic progenitor cells revealed by time-lapse intravital imaging in long bones. Blood. 2009;114:290-298.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 170]  [Cited by in F6Publishing: 179]  [Article Influence: 11.9]  [Reference Citation Analysis (0)]
72.  Sudo K, Ema H, Morita Y, Nakauchi H. Age-associated characteristics of murine hematopoietic stem cells. J Exp Med. 2000;192:1273-1280.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 541]  [Cited by in F6Publishing: 514]  [Article Influence: 21.4]  [Reference Citation Analysis (0)]
73.  Bi S, Wang H, Kuang W. Stem cell rejuvenation and the role of autophagy in age retardation by caloric restriction: An update. Mech Ageing Dev. 2018;175:46-54.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 15]  [Cited by in F6Publishing: 10]  [Article Influence: 1.7]  [Reference Citation Analysis (0)]
74.  Shao L, Li H, Pazhanisamy SK, Meng A, Wang Y, Zhou D. Reactive oxygen species and hematopoietic stem cell senescence. Int J Hematol. 2011;94:24-32.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 124]  [Cited by in F6Publishing: 130]  [Article Influence: 10.0]  [Reference Citation Analysis (0)]
75.  Ermolaeva M, Neri F, Ori A, Rudolph KL. Cellular and epigenetic drivers of stem cell ageing. Nat Rev Mol Cell Biol. 2018;19:594-610.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 126]  [Cited by in F6Publishing: 142]  [Article Influence: 28.4]  [Reference Citation Analysis (0)]
76.  Bradley E, Bieberich E, Mivechi NF, Tangpisuthipongsa D, Wang G. Regulation of embryonic stem cell pluripotency by heat shock protein 90. Stem Cells. 2012;30:1624-1633.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 49]  [Cited by in F6Publishing: 50]  [Article Influence: 4.5]  [Reference Citation Analysis (0)]
77.  Rossi DJ, Bryder D, Seita J, Nussenzweig A, Hoeijmakers J, Weissman IL. Deficiencies in DNA damage repair limit the function of haematopoietic stem cells with age. Nature. 2007;447:725-729.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 808]  [Cited by in F6Publishing: 818]  [Article Influence: 48.1]  [Reference Citation Analysis (0)]
78.  Dykstra B, Olthof S, Schreuder J, Ritsema M, de Haan G. Clonal analysis reveals multiple functional defects of aged murine hematopoietic stem cells. J Exp Med. 2011;208:2691-2703.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 309]  [Cited by in F6Publishing: 317]  [Article Influence: 24.4]  [Reference Citation Analysis (0)]
79.  Liang Y, Van Zant G, Szilvassy SJ. Effects of aging on the homing and engraftment of murine hematopoietic stem and progenitor cells. Blood. 2005;106:1479-1487.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 286]  [Cited by in F6Publishing: 270]  [Article Influence: 14.2]  [Reference Citation Analysis (0)]
80.  Morrison SJ, Wandycz AM, Akashi K, Globerson A, Weissman IL. The aging of hematopoietic stem cells. Nat Med. 1996;2:1011-1016.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 650]  [Cited by in F6Publishing: 674]  [Article Influence: 24.1]  [Reference Citation Analysis (0)]
81.  Beerman I, Bhattacharya D, Zandi S, Sigvardsson M, Weissman IL, Bryder D, Rossi DJ. Functionally distinct hematopoietic stem cells modulate hematopoietic lineage potential during aging by a mechanism of clonal expansion. Proc Natl Acad Sci U S A. 2010;107:5465-5470.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 465]  [Cited by in F6Publishing: 477]  [Article Influence: 34.1]  [Reference Citation Analysis (0)]
82.  Genovese G, Kähler AK, Handsaker RE, Lindberg J, Rose SA, Bakhoum SF, Chambert K, Mick E, Neale BM, Fromer M, Purcell SM, Svantesson O, Landén M, Höglund M, Lehmann S, Gabriel SB, Moran JL, Lander ES, Sullivan PF, Sklar P, Grönberg H, Hultman CM, McCarroll SA. Clonal hematopoiesis and blood-cancer risk inferred from blood DNA sequence. N Engl J Med. 2014;371:2477-2487.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2160]  [Cited by in F6Publishing: 2260]  [Article Influence: 226.0]  [Reference Citation Analysis (0)]
83.  Jaiswal S, Fontanillas P, Flannick J, Manning A, Grauman PV, Mar BG, Lindsley RC, Mermel CH, Burtt N, Chavez A, Higgins JM, Moltchanov V, Kuo FC, Kluk MJ, Henderson B, Kinnunen L, Koistinen HA, Ladenvall C, Getz G, Correa A, Banahan BF, Gabriel S, Kathiresan S, Stringham HM, McCarthy MI, Boehnke M, Tuomilehto J, Haiman C, Groop L, Atzmon G, Wilson JG, Neuberg D, Altshuler D, Ebert BL. Age-related clonal hematopoiesis associated with adverse outcomes. N Engl J Med. 2014;371:2488-2498.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3209]  [Cited by in F6Publishing: 2957]  [Article Influence: 295.7]  [Reference Citation Analysis (0)]
84.  Li Y, Wu Q, Wang Y, Li L, Bu H, Bao J. Senescence of mesenchymal stem cells (Review). Int J Mol Med. 2017;39:775-782.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 155]  [Cited by in F6Publishing: 171]  [Article Influence: 24.4]  [Reference Citation Analysis (0)]
85.  Jiang Y, Mishima H, Sakai S, Liu YK, Ohyabu Y, Uemura T. Gene expression analysis of major lineage-defining factors in human bone marrow cells: effect of aging, gender, and age-related disorders. J Orthop Res. 2008;26:910-917.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 71]  [Cited by in F6Publishing: 71]  [Article Influence: 4.4]  [Reference Citation Analysis (0)]
86.  Xu C, Wang J, Zhu T, Shen Y, Tang X, Fang L, Xu Y. Cross-Talking Between PPAR and WNT Signaling and its Regulation in Mesenchymal Stem Cell Differentiation. Curr Stem Cell Res Ther. 2016;11:247-254.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 55]  [Cited by in F6Publishing: 59]  [Article Influence: 7.4]  [Reference Citation Analysis (0)]
87.  Xie X, Hiona A, Lee AS, Cao F, Huang M, Li Z, Cherry A, Pei X, Wu JC. Effects of long-term culture on human embryonic stem cell aging. Stem Cells Dev. 2011;20:127-138.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 25]  [Cited by in F6Publishing: 27]  [Article Influence: 1.9]  [Reference Citation Analysis (0)]
88.  Nasu A, Ikeya M, Yamamoto T, Watanabe A, Jin Y, Matsumoto Y, Hayakawa K, Amano N, Sato S, Osafune K, Aoyama T, Nakamura T, Kato T, Toguchida J. Genetically matched human iPS cells reveal that propensity for cartilage and bone differentiation differs with clones, not cell type of origin. PLoS One. 2013;8:e53771.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 46]  [Cited by in F6Publishing: 47]  [Article Influence: 4.3]  [Reference Citation Analysis (0)]
89.  Rohani L, Johnson AA, Arnold A, Stolzing A. The aging signature: a hallmark of induced pluripotent stem cells? Aging Cell. 2014;13:2-7.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 62]  [Cited by in F6Publishing: 64]  [Article Influence: 6.4]  [Reference Citation Analysis (0)]
90.  Chambers SM, Shaw CA, Gatza C, Fisk CJ, Donehower LA, Goodell MA. Aging hematopoietic stem cells decline in function and exhibit epigenetic dysregulation. PLoS Biol. 2007;5:e201.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 568]  [Cited by in F6Publishing: 560]  [Article Influence: 35.0]  [Reference Citation Analysis (0)]
91.  Conboy IM, Rando TA. Heterochronic parabiosis for the study of the effects of aging on stem cells and their niches. Cell Cycle. 2012;11:2260-2267.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 152]  [Cited by in F6Publishing: 164]  [Article Influence: 13.7]  [Reference Citation Analysis (0)]
92.  de Haan G, Nijhof W, Van Zant G. Mouse strain-dependent changes in frequency and proliferation of hematopoietic stem cells during aging: correlation between lifespan and cycling activity. Blood. 1997;89:1543-1550.  [PubMed]  [DOI]  [Cited in This Article: ]
93.  Bernet JD, Doles JD, Hall JK, Kelly Tanaka K, Carter TA, Olwin BB. p38 MAPK signaling underlies a cell-autonomous loss of stem cell self-renewal in skeletal muscle of aged mice. Nat Med. 2014;20:265-271.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 372]  [Cited by in F6Publishing: 389]  [Article Influence: 38.9]  [Reference Citation Analysis (0)]
94.  Bortoli S, Renault V, Eveno E, Auffray C, Butler-Browne G, Piétu G. Gene expression profiling of human satellite cells during muscular aging using cDNA arrays. Gene. 2003;321:145-154.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 39]  [Cited by in F6Publishing: 41]  [Article Influence: 2.1]  [Reference Citation Analysis (0)]
95.  Collins CA, Zammit PS, Ruiz AP, Morgan JE, Partridge TA. A population of myogenic stem cells that survives skeletal muscle aging. Stem Cells. 2007;25:885-894.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 234]  [Cited by in F6Publishing: 238]  [Article Influence: 14.0]  [Reference Citation Analysis (0)]
96.  Cosgrove BD, Gilbert PM, Porpiglia E, Mourkioti F, Lee SP, Corbel SY, Llewellyn ME, Delp SL, Blau HM. Rejuvenation of the muscle stem cell population restores strength to injured aged muscles. Nat Med. 2014;20:255-264.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 451]  [Cited by in F6Publishing: 446]  [Article Influence: 44.6]  [Reference Citation Analysis (0)]
97.  Sousa-Victor P, Gutarra S, García-Prat L, Rodriguez-Ubreva J, Ortet L, Ruiz-Bonilla V, Jardí M, Ballestar E, González S, Serrano AL, Perdiguero E, Muñoz-Cánoves P. Geriatric muscle stem cells switch reversible quiescence into senescence. Nature. 2014;506:316-321.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 641]  [Cited by in F6Publishing: 644]  [Article Influence: 64.4]  [Reference Citation Analysis (0)]
98.  Conboy IM, Conboy MJ, Smythe GM, Rando TA. Notch-mediated restoration of regenerative potential to aged muscle. Science. 2003;302:1575-1577.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 801]  [Cited by in F6Publishing: 761]  [Article Influence: 36.2]  [Reference Citation Analysis (0)]
99.  Price FD, von Maltzahn J, Bentzinger CF, Dumont NA, Yin H, Chang NC, Wilson DH, Frenette J, Rudnicki MA. Inhibition of JAK-STAT signaling stimulates adult satellite cell function. Nat Med. 2014;20:1174-1181.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 251]  [Cited by in F6Publishing: 272]  [Article Influence: 27.2]  [Reference Citation Analysis (0)]
100.  HAYFLICK L, MOORHEAD PS. The serial cultivation of human diploid cell strains. Exp Cell Res. 1961;25:585-621.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 5393]  [Cited by in F6Publishing: 5098]  [Article Influence: 196.1]  [Reference Citation Analysis (0)]
101.  Hwang ES. Senescence suppressors: their practical importance in replicative lifespan extension in stem cells. Cell Mol Life Sci. 2014;71:4207-4219.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17]  [Cited by in F6Publishing: 13]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
102.  Baxter MA, Wynn RF, Jowitt SN, Wraith JE, Fairbairn LJ, Bellantuono I. Study of telomere length reveals rapid aging of human marrow stromal cells following in vitro expansion. Stem Cells. 2004;22:675-682.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 571]  [Cited by in F6Publishing: 566]  [Article Influence: 29.8]  [Reference Citation Analysis (0)]
103.  Raz V, Vermolen BJ, Garini Y, Onderwater JJ, Mommaas-Kienhuis MA, Koster AJ, Young IT, Tanke H, Dirks RW. The nuclear lamina promotes telomere aggregation and centromere peripheral localization during senescence of human mesenchymal stem cells. J Cell Sci. 2008;121:4018-4028.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 65]  [Cited by in F6Publishing: 71]  [Article Influence: 4.7]  [Reference Citation Analysis (0)]
104.  Tan FC, Hutchison ER, Eitan E, Mattson MP. Are there roles for brain cell senescence in aging and neurodegenerative disorders? Biogerontology. 2014;15:643-660.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 75]  [Cited by in F6Publishing: 81]  [Article Influence: 8.1]  [Reference Citation Analysis (0)]
105.  Chinta SJ, Woods G, Rane A, Demaria M, Campisi J, Andersen JK. Cellular senescence and the aging brain. Exp Gerontol. 2015;68:3-7.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 164]  [Cited by in F6Publishing: 174]  [Article Influence: 17.4]  [Reference Citation Analysis (0)]
106.  Piechota M, Sunderland P, Wysocka A, Nalberczak M, Sliwinska MA, Radwanska K, Sikora E. Is senescence-associated β-galactosidase a marker of neuronal senescence? Oncotarget. 2016;7:81099-81109.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 64]  [Cited by in F6Publishing: 84]  [Article Influence: 14.0]  [Reference Citation Analysis (0)]
107.  Bernardes de Jesus B, Vera E, Schneeberger K, Tejera AM, Ayuso E, Bosch F, Blasco MA. Telomerase gene therapy in adult and old mice delays aging and increases longevity without increasing cancer. EMBO Mol Med. 2012;4:691-704.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 286]  [Cited by in F6Publishing: 290]  [Article Influence: 24.2]  [Reference Citation Analysis (0)]
108.  Tomás-Loba A, Flores I, Fernández-Marcos PJ, Cayuela ML, Maraver A, Tejera A, Borrás C, Matheu A, Klatt P, Flores JM, Viña J, Serrano M, Blasco MA. Telomerase reverse transcriptase delays aging in cancer-resistant mice. Cell. 2008;135:609-622.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 325]  [Cited by in F6Publishing: 295]  [Article Influence: 18.4]  [Reference Citation Analysis (0)]
109.  Erjavec N, Larsson L, Grantham J, Nyström T. Accelerated aging and failure to segregate damaged proteins in Sir2 mutants can be suppressed by overproducing the protein aggregation-remodeling factor Hsp104p. Genes Dev. 2007;21:2410-2421.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 277]  [Cited by in F6Publishing: 293]  [Article Influence: 17.2]  [Reference Citation Analysis (0)]
110.  Lee HW, Blasco MA, Gottlieb GJ, Horner JW 2nd, Greider CW, DePinho EA. Essential role of mouse telomerase in highly proliferative organs. Nature. 1998;392:569-574.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 963]  [Cited by in F6Publishing: 943]  [Article Influence: 36.3]  [Reference Citation Analysis (0)]
111.  Kanaar R, Wyman C, Rothstein R. Quality control of DNA break metabolism: in the 'end', it's a good thing. EMBO J. 2008;27:581-588.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 48]  [Cited by in F6Publishing: 52]  [Article Influence: 3.3]  [Reference Citation Analysis (0)]
112.  Walter D, Lier A, Geiselhart A, Thalheimer FB, Huntscha S, Sobotta MC, Moehrle B, Brocks D, Bayindir I, Kaschutnig P, Muedder K, Klein C, Jauch A, Schroeder T, Geiger H, Dick TP, Holland-Letz T, Schmezer P, Lane SW, Rieger MA, Essers MA, Williams DA, Trumpp A, Milsom MD. Exit from dormancy provokes DNA-damage-induced attrition in haematopoietic stem cells. Nature. 2015;520:549-552.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 392]  [Cited by in F6Publishing: 438]  [Article Influence: 48.7]  [Reference Citation Analysis (0)]
113.  Lamming DW, Bar-Peled L. Lysosome: The metabolic signaling hub. Traffic. 2019;20:27-38.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 65]  [Cited by in F6Publishing: 86]  [Article Influence: 14.3]  [Reference Citation Analysis (0)]
114.  Szymańska P, Martin KR, MacKeigan JP, Hlavacek WS, Lipniacki T. Computational analysis of an autophagy/translation switch based on mutual inhibition of MTORC1 and ULK1. PLoS One. 2015;10:e0116550.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 35]  [Cited by in F6Publishing: 31]  [Article Influence: 3.4]  [Reference Citation Analysis (0)]
115.  Revuelta M, Matheu A. Autophagy in stem cell aging. Aging Cell. 2017;16:912-915.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 74]  [Cited by in F6Publishing: 93]  [Article Influence: 13.3]  [Reference Citation Analysis (0)]
116.  Cuervo AM. Autophagy and aging: keeping that old broom working. Trends Genet. 2008;24:604-612.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 432]  [Cited by in F6Publishing: 407]  [Article Influence: 25.4]  [Reference Citation Analysis (0)]
117.  Rubinsztein DC, Mariño G, Kroemer G. Autophagy and aging. Cell. 2011;146:682-695.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1509]  [Cited by in F6Publishing: 1533]  [Article Influence: 117.9]  [Reference Citation Analysis (0)]
118.  Bufalino MR, DeVeale B, van der Kooy D. The asymmetric segregation of damaged proteins is stem cell-type dependent. J Cell Biol. 2013;201:523-530.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 76]  [Cited by in F6Publishing: 70]  [Article Influence: 6.4]  [Reference Citation Analysis (0)]
119.  Katajisto P, Döhla J, Chaffer CL, Pentinmikko N, Marjanovic N, Iqbal S, Zoncu R, Chen W, Weinberg RA, Sabatini DM. Stem cells. Asymmetric apportioning of aged mitochondria between daughter cells is required for stemness. Science. 2015;348:340-343.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 359]  [Cited by in F6Publishing: 368]  [Article Influence: 40.9]  [Reference Citation Analysis (0)]
120.  Rujano MA, Bosveld F, Salomons FA, Dijk F, van Waarde MA, van der Want JJ, de Vos RA, Brunt ER, Sibon OC, Kampinga HH. Polarised asymmetric inheritance of accumulated protein damage in higher eukaryotes. PLoS Biol. 2006;4:e417.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 183]  [Cited by in F6Publishing: 187]  [Article Influence: 11.0]  [Reference Citation Analysis (0)]
121.  Salemi S, Yousefi S, Constantinescu MA, Fey MF, Simon HU. Autophagy is required for self-renewal and differentiation of adult human stem cells. Cell Res. 2012;22:432-435.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 136]  [Cited by in F6Publishing: 132]  [Article Influence: 10.2]  [Reference Citation Analysis (0)]
122.  García-Prat L, Martínez-Vicente M, Perdiguero E, Ortet L, Rodríguez-Ubreva J, Rebollo E, Ruiz-Bonilla V, Gutarra S, Ballestar E, Serrano AL, Sandri M, Muñoz-Cánoves P. Autophagy maintains stemness by preventing senescence. Nature. 2016;529:37-42.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 791]  [Cited by in F6Publishing: 882]  [Article Influence: 110.3]  [Reference Citation Analysis (0)]
123.  Ho TT, Warr MR, Adelman ER, Lansinger OM, Flach J, Verovskaya EV, Figueroa ME, Passegué E. Autophagy maintains the metabolism and function of young and old stem cells. Nature. 2017;543:205-210.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 490]  [Cited by in F6Publishing: 572]  [Article Influence: 81.7]  [Reference Citation Analysis (0)]
124.  Iqbal K, Jin SG, Pfeifer GP, Szabó PE. Reprogramming of the paternal genome upon fertilization involves genome-wide oxidation of 5-methylcytosine. Proc Natl Acad Sci U S A. 2011;108:3642-3647.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 533]  [Cited by in F6Publishing: 510]  [Article Influence: 39.2]  [Reference Citation Analysis (0)]
125.  Rideout WM, Eggan K, Jaenisch R. Nuclear cloning and epigenetic reprogramming of the genome. Science. 2001;293:1093-1098.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 570]  [Cited by in F6Publishing: 586]  [Article Influence: 25.5]  [Reference Citation Analysis (0)]
126.  Rando TA, Chang HY. Aging, rejuvenation, and epigenetic reprogramming: resetting the aging clock. Cell. 2012;148:46-57.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 344]  [Cited by in F6Publishing: 330]  [Article Influence: 27.5]  [Reference Citation Analysis (0)]
127.  Liang G, Zhang Y. Genetic and epigenetic variations in iPSCs: potential causes and implications for application. Cell Stem Cell. 2013;13:149-159.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 264]  [Cited by in F6Publishing: 270]  [Article Influence: 27.0]  [Reference Citation Analysis (0)]
128.  Wahlestedt M, Norddahl GL, Sten G, Ugale A, Frisk MA, Mattsson R, Deierborg T, Sigvardsson M, Bryder D. An epigenetic component of hematopoietic stem cell aging amenable to reprogramming into a young state. Blood. 2013;121:4257-4264.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 74]  [Cited by in F6Publishing: 72]  [Article Influence: 6.5]  [Reference Citation Analysis (0)]
129.  Frobel J, Hemeda H, Lenz M, Abagnale G, Joussen S, Denecke B, Sarić T, Zenke M, Wagner W. Epigenetic rejuvenation of mesenchymal stromal cells derived from induced pluripotent stem cells. Stem Cell Reports. 2014;3:414-422.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 152]  [Cited by in F6Publishing: 163]  [Article Influence: 16.3]  [Reference Citation Analysis (0)]
130.  Abad M, Mosteiro L, Pantoja C, Cañamero M, Rayon T, Ors I, Graña O, Megías D, Domínguez O, Martínez D, Manzanares M, Ortega S, Serrano M. Reprogramming in vivo produces teratomas and iPS cells with totipotency features. Nature. 2013;502:340-345.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 336]  [Cited by in F6Publishing: 349]  [Article Influence: 31.7]  [Reference Citation Analysis (0)]
131.  Mertens J, Paquola ACM, Ku M, Hatch E, Böhnke L, Ladjevardi S, McGrath S, Campbell B, Lee H, Herdy JR, Gonçalves JT, Toda T, Kim Y, Winkler J, Yao J, Hetzer MW, Gage FH. Directly Reprogrammed Human Neurons Retain Aging-Associated Transcriptomic Signatures and Reveal Age-Related Nucleocytoplasmic Defects. Cell Stem Cell. 2015;17:705-718.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 428]  [Cited by in F6Publishing: 472]  [Article Influence: 52.4]  [Reference Citation Analysis (0)]
132.  Hackl M, Brunner S, Fortschegger K, Schreiner C, Micutkova L, Mück C, Laschober GT, Lepperdinger G, Sampson N, Berger P, Herndler-Brandstetter D, Wieser M, Kühnel H, Strasser A, Rinnerthaler M, Breitenbach M, Mildner M, Eckhart L, Tschachler E, Trost A, Bauer JW, Papak C, Trajanoski Z, Scheideler M, Grillari-Voglauer R, Grubeck-Loebenstein B, Jansen-Dürr P, Grillari J. miR-17, miR-19b, miR-20a, and miR-106a are down-regulated in human aging. Aging Cell. 2010;9:291-296.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 275]  [Cited by in F6Publishing: 279]  [Article Influence: 19.9]  [Reference Citation Analysis (0)]
133.  Okada M, Kim HW, Matsu-ura K, Wang YG, Xu M, Ashraf M. Abrogation of Age-Induced MicroRNA-195 Rejuvenates the Senescent Mesenchymal Stem Cells by Reactivating Telomerase. Stem Cells. 2016;34:148-159.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 64]  [Cited by in F6Publishing: 79]  [Article Influence: 8.8]  [Reference Citation Analysis (0)]
134.  Lin RC, Weeks KL, Gao XM, Williams RB, Bernardo BC, Kiriazis H, Matthews VB, Woodcock EA, Bouwman RD, Mollica JP, Speirs HJ, Dawes IW, Daly RJ, Shioi T, Izumo S, Febbraio MA, Du XJ, McMullen JR. PI3K(p110 alpha) protects against myocardial infarction-induced heart failure: identification of PI3K-regulated miRNA and mRNA. Arterioscler Thromb Vasc Biol. 2010;30:724-732.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 129]  [Cited by in F6Publishing: 135]  [Article Influence: 9.6]  [Reference Citation Analysis (0)]
135.  Zhang F, Cui J, Liu X, Lv B, Liu X, Xie Z, Yu B. Roles of microRNA-34a targeting SIRT1 in mesenchymal stem cells. Stem Cell Res Ther. 2015;6:195.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 56]  [Cited by in F6Publishing: 64]  [Article Influence: 7.1]  [Reference Citation Analysis (0)]
136.  Lee S, Jung JW, Park SB, Roh K, Lee SY, Kim JH, Kang SK, Kang KS. Histone deacetylase regulates high mobility group A2-targeting microRNAs in human cord blood-derived multipotent stem cell aging. Cell Mol Life Sci. 2011;68:325-336.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 75]  [Cited by in F6Publishing: 77]  [Article Influence: 5.5]  [Reference Citation Analysis (0)]
137.  So AY, Jung JW, Lee S, Kim HS, Kang KS. DNA methyltransferase controls stem cell aging by regulating BMI1 and EZH2 through microRNAs. PLoS One. 2011;6:e19503.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 136]  [Cited by in F6Publishing: 136]  [Article Influence: 10.5]  [Reference Citation Analysis (0)]
138.  Hodzic M, Naaldijk Y, Stolzing A. Regulating aging in adult stem cells with microRNA. Z Gerontol Geriatr. 2013;46:629-634.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 18]  [Cited by in F6Publishing: 15]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
139.  Siddiqi S, Sussman MA. Cardiac Hegemony of Senescence. Curr Transl Geriatr Exp Gerontol Rep. 2013;2.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 15]  [Cited by in F6Publishing: 15]  [Article Influence: 1.4]  [Reference Citation Analysis (0)]
140.  Liang R, Ghaffari S. Stem cells, redox signaling, and stem cell aging. Antioxid Redox Signal. 2014;20:1902-1916.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 75]  [Cited by in F6Publishing: 73]  [Article Influence: 7.3]  [Reference Citation Analysis (0)]
141.  Simsek T, Kocabas F, Zheng J, Deberardinis RJ, Mahmoud AI, Olson EN, Schneider JW, Zhang CC, Sadek HA. The distinct metabolic profile of hematopoietic stem cells reflects their location in a hypoxic niche. Cell Stem Cell. 2010;7:380-390.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 737]  [Cited by in F6Publishing: 788]  [Article Influence: 56.3]  [Reference Citation Analysis (0)]
142.  Ok JS, Song SB, Hwang ES. Enhancement of Replication and Differentiation Potential of Human Bone Marrow Stem Cells by Nicotinamide Treatment. Int J Stem Cells. 2018;11:13-25.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 11]  [Cited by in F6Publishing: 17]  [Article Influence: 2.8]  [Reference Citation Analysis (0)]
143.  Kang HT, Lee HI, Hwang ES. Nicotinamide extends replicative lifespan of human cells. Aging Cell. 2006;5:423-436.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 67]  [Cited by in F6Publishing: 70]  [Article Influence: 3.9]  [Reference Citation Analysis (0)]
144.  Mouchiroud L, Houtkooper RH, Moullan N, Katsyuba E, Ryu D, Cantó C, Mottis A, Jo YS, Viswanathan M, Schoonjans K, Guarente L, Auwerx J. The NAD(+)/Sirtuin Pathway Modulates Longevity through Activation of Mitochondrial UPR and FOXO Signaling. Cell. 2013;154:430-441.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 773]  [Cited by in F6Publishing: 833]  [Article Influence: 75.7]  [Reference Citation Analysis (0)]
145.  Roos CM, Hagler M, Zhang B, Oehler EA, Arghami A, Miller JD. Transcriptional and phenotypic changes in aorta and aortic valve with aging and MnSOD deficiency in mice. Am J Physiol Heart Circ Physiol. 2013;305:H1428-H1439.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 42]  [Cited by in F6Publishing: 38]  [Article Influence: 3.5]  [Reference Citation Analysis (0)]
146.  Wang XQ, Shao Y, Ma CY, Chen W, Sun L, Liu W, Zhang DY, Fu BC, Liu KY, Jia ZB, Xie BD, Jiang SL, Li RK, Tian H. Decreased SIRT3 in aged human mesenchymal stromal/stem cells increases cellular susceptibility to oxidative stress. J Cell Mol Med. 2014;18:2298-2310.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 42]  [Cited by in F6Publishing: 43]  [Article Influence: 4.3]  [Reference Citation Analysis (0)]
147.  Zhai XY, Yan P, Zhang J, Song HF, Yin WJ, Gong H, Li H, Wu J, Xie J, Li RK. Knockdown of SIRT6 Enables Human Bone Marrow Mesenchymal Stem Cell Senescence. Rejuvenation Res. 2016;19:373-384.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 15]  [Cited by in F6Publishing: 19]  [Article Influence: 2.4]  [Reference Citation Analysis (0)]
148.  Moore D. L., Pilz G. A., Araúzo-Bravo M. J., Barral Y., Jessberger S.. A mechanism for the segregation of age in mammalian neural stem cells. Science. 2015;349:1334-1338.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 110]  [Cited by in F6Publishing: 112]  [Article Influence: 12.4]  [Reference Citation Analysis (0)]
149.  Paneni F, Diaz Cañestro C, Libby P, Lüscher TF, Camici GG. The Aging Cardiovascular System: Understanding It at the Cellular and Clinical Levels. J Am Coll Cardiol. 2017;69:1952-1967.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 288]  [Cited by in F6Publishing: 304]  [Article Influence: 43.4]  [Reference Citation Analysis (0)]
150.  Petrini S, Borghi R, D'Oria V, Restaldi F, Moreno S, Novelli A, Bertini E, Compagnucci C. Aged induced pluripotent stem cell (iPSCs) as a new cellular model for studying premature aging. Aging (Albany NY). 2017;9:1453-1469.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 26]  [Reference Citation Analysis (0)]
151.  Wątroba M, Szukiewicz D. The role of sirtuins in aging and age-related diseases. Adv Med Sci. 2016;61:52-62.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 110]  [Cited by in F6Publishing: 99]  [Article Influence: 12.4]  [Reference Citation Analysis (0)]
152.  Yao H, Rahman I. Perspectives on translational and therapeutic aspects of SIRT1 in inflammaging and senescence. Biochem Pharmacol. 2012;84:1332-1339.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 86]  [Cited by in F6Publishing: 93]  [Article Influence: 7.8]  [Reference Citation Analysis (0)]
153.  Zhu C, Hu W, Wu H, Hu X. No evident dose-response relationship between cellular ROS level and its cytotoxicity--a paradoxical issue in ROS-based cancer therapy. Sci Rep. 2014;4:5029.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 78]  [Cited by in F6Publishing: 73]  [Article Influence: 7.3]  [Reference Citation Analysis (0)]
154.  Le Gal K, Ibrahim MX, Wiel C, Sayin VI, Akula MK, Karlsson C, Dalin MG, Akyürek LM, Lindahl P, Nilsson J, Bergo MO. Antioxidants can increase melanoma metastasis in mice. Sci Transl Med. 2015;7:308re8.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 373]  [Cited by in F6Publishing: 401]  [Article Influence: 44.6]  [Reference Citation Analysis (0)]
155.  Biesalski HK, Grune T, Tinz J, Zöllner I, Blumberg JB. Reexamination of a meta-analysis of the effect of antioxidant supplementation on mortality and health in randomized trials. Nutrients. 2010;2:929-949.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 47]  [Cited by in F6Publishing: 49]  [Article Influence: 3.5]  [Reference Citation Analysis (0)]
156.  D'Aniello C, Cermola F, Patriarca EJ, Minchiotti G. Vitamin C in Stem Cell Biology: Impact on Extracellular Matrix Homeostasis and Epigenetics. Stem Cells Int. 2017;2017:8936156.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 53]  [Cited by in F6Publishing: 57]  [Article Influence: 8.1]  [Reference Citation Analysis (0)]
157.  Gertz M, Nguyen GT, Fischer F, Suenkel B, Schlicker C, Fränzel B, Tomaschewski J, Aladini F, Becker C, Wolters D, Steegborn C. A molecular mechanism for direct sirtuin activation by resveratrol. PLoS One. 2012;7:e49761.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 186]  [Cited by in F6Publishing: 196]  [Article Influence: 16.3]  [Reference Citation Analysis (0)]
158.  Imai S, Guarente L. NAD+ and sirtuins in aging and disease. Trends Cell Biol. 2014;24:464-471.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 720]  [Cited by in F6Publishing: 864]  [Article Influence: 86.4]  [Reference Citation Analysis (0)]
159.  Imai SI, Guarente L. It takes two to tango: NAD+ and sirtuins in aging/longevity control. NPJ Aging Mech Dis. 2016;2:16017.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 218]  [Cited by in F6Publishing: 238]  [Article Influence: 29.8]  [Reference Citation Analysis (0)]
160.  Chao SC, Chen YJ, Huang KH, Kuo KL, Yang TH, Huang KY, Wang CC, Tang CH, Yang RS, Liu SH. Induction of sirtuin-1 signaling by resveratrol induces human chondrosarcoma cell apoptosis and exhibits antitumor activity. Sci Rep. 2017;7:3180.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 52]  [Cited by in F6Publishing: 53]  [Article Influence: 7.6]  [Reference Citation Analysis (0)]
161.  Zhu Y, Tchkonia T, Pirtskhalava T, Gower AC, Ding H, Giorgadze N, Palmer AK, Ikeno Y, Hubbard GB, Lenburg M, O'Hara SP, LaRusso NF, Miller JD, Roos CM, Verzosa GC, LeBrasseur NK, Wren JD, Farr JN, Khosla S, Stout MB, McGowan SJ, Fuhrmann-Stroissnigg H, Gurkar AU, Zhao J, Colangelo D, Dorronsoro A, Ling YY, Barghouthy AS, Navarro DC, Sano T, Robbins PD, Niedernhofer LJ, Kirkland JL. The Achilles' heel of senescent cells: from transcriptome to senolytic drugs. Aging Cell. 2015;14:644-658.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1102]  [Cited by in F6Publishing: 1379]  [Article Influence: 153.2]  [Reference Citation Analysis (0)]
162.  Korotchkina LG, Leontieva OV, Bukreeva EI, Demidenko ZN, Gudkov AV, Blagosklonny MV. The choice between p53-induced senescence and quiescence is determined in part by the mTOR pathway. Aging (Albany NY). 2010;2:344-352.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 185]  [Cited by in F6Publishing: 223]  [Article Influence: 15.9]  [Reference Citation Analysis (0)]
163.  Anisimov VN. Metformin for aging and cancer prevention. Aging (Albany NY). 2010;2:760-774.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 108]  [Cited by in F6Publishing: 104]  [Article Influence: 7.4]  [Reference Citation Analysis (0)]
164.  Anisimov VN. Metformin: do we finally have an anti-aging drug? Cell Cycle. 2013;12:3483-3489.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 84]  [Cited by in F6Publishing: 83]  [Article Influence: 7.5]  [Reference Citation Analysis (0)]
165.  Madeo F, Tavernarakis N, Kroemer G. Can autophagy promote longevity? Nat Cell Biol. 2010;12:842-846.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 334]  [Cited by in F6Publishing: 354]  [Article Influence: 25.3]  [Reference Citation Analysis (0)]
166.  Madeo F, Zimmermann A, Maiuri MC, Kroemer G. Essential role for autophagy in life span extension. J Clin Invest. 2015;125:85-93.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 336]  [Cited by in F6Publishing: 301]  [Article Influence: 33.4]  [Reference Citation Analysis (0)]
167.  López-Otín C, Blasco MA, Partridge L, Serrano M, Kroemer G. The hallmarks of aging. Cell. 2013;153:1194-1217.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 7855]  [Cited by in F6Publishing: 8856]  [Article Influence: 805.1]  [Reference Citation Analysis (0)]
168.  Avolio E, Gianfranceschi G, Cesselli D, Caragnano A, Athanasakis E, Katare R, Meloni M, Palma A, Barchiesi A, Vascotto C, Toffoletto B, Mazzega E, Finato N, Aresu G, Livi U, Emanueli C, Scoles G, Beltrami CA, Madeddu P, Beltrami AP. Ex vivo molecular rejuvenation improves the therapeutic activity of senescent human cardiac stem cells in a mouse model of myocardial infarction. Stem Cells. 2014;32:2373-2385.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 53]  [Cited by in F6Publishing: 55]  [Article Influence: 6.1]  [Reference Citation Analysis (0)]
169.  Cerletti M, Jang YC, Finley LW, Haigis MC, Wagers AJ. Short-term calorie restriction enhances skeletal muscle stem cell function. Cell Stem Cell. 2012;10:515-519.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 271]  [Cited by in F6Publishing: 265]  [Article Influence: 22.1]  [Reference Citation Analysis (0)]
170.  Mihaylova MM, Sabatini DM, Yilmaz H. Dietary and metabolic control of stem cell function in physiology and cancer. Cell Stem Cell. 2014;14:292-305.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 114]  [Cited by in F6Publishing: 112]  [Article Influence: 11.2]  [Reference Citation Analysis (0)]
171.  Balasubramanian P, Howell PR, Anderson RM. Aging and Caloric Restriction Research: A Biological Perspective With Translational Potential. EBioMedicine. 2017;21:37-44.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 97]  [Cited by in F6Publishing: 93]  [Article Influence: 13.3]  [Reference Citation Analysis (0)]
172.  Bishop NA, Guarente L. Genetic links between diet and lifespan: shared mechanisms from yeast to humans. Nat Rev Genet. 2007;8:835-844.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 318]  [Cited by in F6Publishing: 312]  [Article Influence: 18.4]  [Reference Citation Analysis (0)]
173.  Fontana L, Partridge L, Longo VD. Extending healthy life span--from yeast to humans. Science. 2010;328:321-326.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2226]  [Cited by in F6Publishing: 2008]  [Article Influence: 143.4]  [Reference Citation Analysis (0)]
174.  Ingram DK, Roth GS. Calorie restriction mimetics: can you have your cake and eat it, too? Ageing Res Rev. 2015;20:46-62.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 106]  [Cited by in F6Publishing: 92]  [Article Influence: 10.2]  [Reference Citation Analysis (0)]
175.  Mattison JA, Colman RJ, Beasley TM, Allison DB, Kemnitz JW, Roth GS, Ingram DK, Weindruch R, de Cabo R, Anderson RM. Caloric restriction improves health and survival of rhesus monkeys. Nat Commun. 2017;8:14063.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 470]  [Cited by in F6Publishing: 510]  [Article Influence: 72.9]  [Reference Citation Analysis (0)]
176.  Rusli F, Boekschoten MV, Borelli V, Sun C, Lute C, Menke AL, van den Heuvel J, Salvioli S, Franceschi C, Müller M, Steegenga WT. Plasticity of lifelong calorie-restricted C57BL/6J mice in adapting to a medium-fat diet intervention at old age. Aging Cell. 2018;17.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 4]  [Cited by in F6Publishing: 6]  [Article Influence: 1.0]  [Reference Citation Analysis (0)]
177.  Na HJ, Park JS, Pyo JH, Lee SH, Jeon HJ, Kim YS, Yoo MA. Mechanism of metformin: inhibition of DNA damage and proliferative activity in Drosophila midgut stem cell. Mech Ageing Dev. 2013;134:381-390.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 44]  [Cited by in F6Publishing: 42]  [Article Influence: 3.8]  [Reference Citation Analysis (0)]
178.  Ramot Y, Tiede S, Bíró T, Abu Bakar MH, Sugawara K, Philpott MP, Harrison W, Pietilä M, Paus R. Spermidine promotes human hair growth and is a novel modulator of human epithelial stem cell functions. PLoS One. 2011;6:e22564.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 51]  [Cited by in F6Publishing: 53]  [Article Influence: 4.1]  [Reference Citation Analysis (0)]
179.  Petkovich DA, Podolskiy DI, Lobanov AV, Lee SG, Miller RA, Gladyshev VN. Using DNA Methylation Profiling to Evaluate Biological Age and Longevity Interventions. Cell Metab. 2017;25:954-960.e6.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 251]  [Cited by in F6Publishing: 206]  [Article Influence: 29.4]  [Reference Citation Analysis (0)]
180.  Takeuchi M, Takeuchi K, Kohara A, Satoh M, Shioda S, Ozawa Y, Ohtani A, Morita K, Hirano T, Terai M, Umezawa A, Mizusawa H. Chromosomal instability in human mesenchymal stem cells immortalized with human papilloma virus E6, E7, and hTERT genes. In Vitro Cell Dev Biol Anim. 2007;43:129-138.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 34]  [Cited by in F6Publishing: 37]  [Article Influence: 2.2]  [Reference Citation Analysis (0)]
181.  Simonsen JL, Rosada C, Serakinci N, Justesen J, Stenderup K, Rattan SI, Jensen TG, Kassem M. Telomerase expression extends the proliferative life-span and maintains the osteogenic potential of human bone marrow stromal cells. Nat Biotechnol. 2002;20:592-596.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 616]  [Cited by in F6Publishing: 600]  [Article Influence: 27.3]  [Reference Citation Analysis (0)]
182.  Galderisi U, Cipollaro M, Giordano A. The retinoblastoma gene is involved in multiple aspects of stem cell biology. Oncogene. 2006;25:5250-5256.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 51]  [Cited by in F6Publishing: 58]  [Article Influence: 3.2]  [Reference Citation Analysis (0)]
183.  Gu Z, Cao X, Jiang J, Li L, Da Z, Liu H, Cheng C. Upregulation of p16INK4A promotes cellular senescence of bone marrow-derived mesenchymal stem cells from systemic lupus erythematosus patients. Cell Signal. 2012;24:2307-2314.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 52]  [Cited by in F6Publishing: 42]  [Article Influence: 3.5]  [Reference Citation Analysis (0)]