Skip to main content

Epidermal stem cells: skin surveillance and clinical perspective

Abstract

The skin epidermis is continually influenced by a myriad of internal and external elements. At its basal layer reside epidermal stem cells, which fuels epidermal renovation and hair regeneration with powerful self-renewal ability, as well as keeping diverse signals that direct their activity under surveillance with quick response. The importance of epidermal stem cells in wound healing and immune-related skin conditions has been increasingly recognized, and their potential for clinical applications is attracting attention. In this review, we delve into recent advancements and the various physiological and psychological factors that govern distinct epidermal stem cell populations, including psychological stress, mechanical forces, chronic aging, and circadian rhythm, as well as providing an overview of current methodological approaches. Furthermore, we discuss the pathogenic role of epidermal stem cells in immune-related skin disorders and their potential clinical applications.

Background

The skin serves as a vital barrier for mammals, providing protection against injuries, infections, radiation, and water loss. It also plays a key role in regulating temperature and enabling sensory perception [1, 2]. Structurally, the skin comprises two primary layers including the outer epidermis and the inner dermis as illustrated in Fig. 1. The epidermis, is a multilayered stratified epithelium characterized by a high turnover rate, primarily due to the continuous shedding of its superficial cornified cells [3]. This epidermal layer encompasses the interfollicular epidermis along with various functional appendages, including hair follicles (HFs), sebaceous glands, and sweat glands, each playing distinct roles [4]. Notably, the thickness of the interfollicular epidermis and the distribution of these appendages exhibit considerable variation across different regions of the body [5].

Fig. 1
figure 1

Architecture of the skin and specific display of the interfollicular epidermis. a Skin Composition: the skin epidermis includes the interfollicular epidermis, extending to hair follicle infundibula, and other structures like sebaceous and sweat glands. The dermis, separated from the epidermis by a basement membrane, contains arrector pili muscles, nerves, blood vessels, cells, and extracellular matrix. Below this is the hypodermis, made up of dermal white adipose tissue. b Epidermal differentiation models: in the stochastic model, basal cells are uniform, and each division randomly results in: (i) one dividing progenitor and one differentiating cell moving upwards, (ii) two differentiating cells, or (iii) two progenitors. Conversely, the hierarchical model involves limited stem cells sporadically dividing into transit amplifying cells, which rapidly divide into numerous differentiated cells. As these cells move towards the surface, they differentiate, forming layers: basal, spinous, granular, and the continuously shedding stratum corneum

As the engine of the epidermis, epidermal stem cells (EpSCs) are located in the basal layer of the skin, imparting the skin with remarkable regenerative potential [6]. Not only can they fuel the epidermal turnover and hair regeneration during homeostasis, but they can also be activated to heal the wounds upon injuries [6]. However, dysfunctional EpSCs serve as malignant roots of skin disorders [7]. In this review, we primarily highlight recent advancements in various EpSC populations and their niche. We subsequently discuss multiple factors that exert influence on EpSCs and elucidate their significance, as well as summarizing classic signaling pathways and related metabolism. We also outline methods for studying EpSCs, ranging from in vitro experiments to in vivo approaches, which may inspire future researchers with practicable tools.

Epidermal stem cells in maintaining skin homeostasis

EpSCs universally exhibit the essential ability for self-renewal and differentiation, which can be categorized based on their specific locations into: interfollicular epidermal stem cells (IFESCs), HF stem cells (HFSCs), and sweat gland stem cells. A range of biomarkers are employed to identify these distinct EpSC clusters, each with unique functional properties (Fig. 2).

Fig. 2
figure 2

Architecture of the hair follicle and cluster diversity of epidermal stem cells. This illustration uses a color-coded legend to identify and map distinct stem cell and differentiated cell clusters across various regions of the hair follicle and skin. These regions include the hair follicle infundibulum, junctional zone, isthmus, bulge, hair germ, sebaceous gland and duct, and the interfollicular epidermis

Stem cells in the interfollicular epidermis

The interfollicular epidermis serves as a crucial lipid barrier and is continuously renewed by IFESCs on the basal layers [8]. To maintain a stable stem cell pool and produce differentiated cells, IFESCs employ both symmetric and asymmetric divisions. Symmetric division, aligned with the basal membrane axis, yields two identical daughter cells and guarantees rapid expansion of the epidermis [9]. In contrast, asymmetric division, perpendicular to the basal membrane axis, generates daughter cells with different phenotypes, crucial for epithelial stratification [10]. Interestingly, recent studies suggest that perpendicular division of basal progenitors does not always result in distinct fates for daughter cells, with fate determination depending on separation from the basal layer [11].

Exploring how IFESCs facilitate epidermal turnover via symmetric and asymmetric divisions has given rise to various models. For over a decade, the stochastic and hierarchical models have been central to this debate [6]. The stochastic model suggests that the random self-renewal and differentiation of homogeneous IFESCs lead to basal and committed daughter cells [12], whereas the hierarchical model posits that rare IFESC divisions produce transit-amplifying cells (TACs), which subsequently generate differentiated cells in the upper layers [13]. Recent insights, however, indicate a more nuanced process, transitioning gradually from IFESCs to differentiated keratinocytes. Single-cell transcriptome studies have revealed four spatially distinct stem cell populations in the basal layer of both murine [14] and human epidermis [15], leading to a refined 'hierarchical-lineage' model of epidermal homeostasis [16]. This model acknowledges multiple stem and progenitor cell states [17]. Intravital imaging and a live cell reporter for the Keratin10 (Krt10) gene have tracked progressive transcriptional changes in differentiating IFESCs [11, 18], suggesting that the diversity in molecular profiles of proliferating basal cells can explain observed variations in stem cell commitment. Further research is essential to fully understand the cellular and molecular mechanisms at play.

Stem cells in the hair follicle

The HF operates as a mini-organ, cyclically transitioning between involution and regeneration throughout life, crucial for hair production and body temperature regulation [19]. Structurally, the HF is a cylindrical assembly of epithelial keratinocytes and a fibroblast-rich dermal papilla (Fig. 2), segmented into the infundibulum, isthmus, and lower follicle connecting to the sebaceous gland [20]. The bulge region, located below the isthmus and extending into the lower follicle, is a critical site housing the hair germ and dermal papilla [21]. HFSCs, initially identified as slow-cycling 'label-retaining' cells in the bulge, are now known to inhabit all HF segments [4]. HFSCs from the bulge and isthmus exhibit high clonogenicity, capable of generating various lineages including the interfollicular epidermis, HFs, and sebaceous glands following transplantation or injury [22, 23]. A recent 'telescope model' of HF development has been proposed, illustrating the concentric arrangement of epithelial lineage precursors in the basal layer of the hair placode. These precursors extend towards the dermis to form a cylindrical structure with longitudinally aligned compartments [24].

The hair cycle's completion hinges on the orchestrated actions of HFSCs. It includes prolonged hair growth (anagen), brief degeneration (catagen), and a resting phase (telogen) [8]. In the telogen phase, HFSCs in both the bulge and hair germ are dormant. With the anagen phase's onset, hair germ stem cells rapidly proliferate, forming a matrix of TACs that divide swiftly to construct the hair shaft and inner root sheath. Bulge stem cells later become active, contributing to the outer root sheath. During catagen, matrix and inner root sheath cells undergo apoptosis [2]. Meanwhile, some lower outer root sheath cells differentiate into inner Krt6+ bulge cells, middle cells initiate a new hair germ, and upper cells form a new bulge, setting the stage for the next hair cycle [19]. Notably, recent single-cell RNA sequencing studies on human scalp HFs have uncovered significant differences from mouse models, such as the presence of epithelial-mesenchymal transition features uniquely in human HFSCs [25].

Melanocyte stem cells (MeSCs), originating from the neural crest, notably share the bulge niche with HFSCs, contributing to hair pigmentation [26]. These MeSCs differentiate into melanocytes that migrate to the hair matrix during the onset of each anagen phase and undergo apoptosis with matrix TACs in the catagen phase [25, 26]. Recent studies have revealed that MeSC maintenance involves not only resident stem cells but also those reverting from a transit-amplifying state [27]. Furthermore, senescent melanocytes secreting Osteopontin have been found to stimulate adjacent HFSC activity and hair regeneration, shedding light on the dynamic interplay between EpSCs and melanocytes [28].

Stem cells in the sweat gland

Sweat glands, ubiquitous secretory epidermal appendages, are categorized mainly into eccrine and apocrine types in humans, with eccrine glands constituting about 90% of the total but notably absent in most lab animal models [3, 29]. Eccrine glands feature a coiled secretory section in the dermis, composed of myoepithelial and luminal cells, and a ductal portion extending to the epidermis, lined by two to three layers of epithelial cells [29]. Research on sweat gland stem/progenitor cells remains in its nascent stages. To date, two progenitor populations have been identified in developing sweat ducts: multipotent Krt14+ progenitors forming myoepithelial cells and transient, proliferative Krt14low/Krt18+ suprabasal progenitors generating luminal cells, both displaying multipotency during morphogenesis [30]. In adults, four distinct stem cell populations in sweat glands exhibit diverse behaviors in contexts like homeostasis, wound repair, and transplantation. Typically, sweat ducts undergo significant homeostatic turnover, akin to the epidermis, while the glandular portion remains relatively quiescent [30].

Diverse cell types interacting with epidermal stem cells in niche

The niche where stem cells reside, was traditionally understood to comprise heterologous cell populations like fibroblasts, immune cells, sensory neurons, arrector pili muscles, and blood vessels [19]. However, recent findings have identified apoptotic EpSCs as a novel niche component, significantly influencing stem cell proliferation and tissue regeneration, primarily through Wnt3 release [31].

Interactions between stem cells and their niches are bidirectional [32]. Not only does the niche influence stem cells and progenitors, but these cells also actively shape their residing niche [33]. Recent discoveries include the dermal sheath's role as smooth muscle, generating a constrictive force to position the dermal papilla near stem cells, vital for hair growth activation [34]. Further studies have revealed that this contraction is initiated by epithelial progenitors in the outer root sheath via endothelin signaling [35]. Additionally, the sympathetic neuron and arrector pili muscles, attached to HFs, form a dual-component niche regulating HF regeneration in adults. This system operates through norepinephrine released by sympathetic nerves, with the arrector pili muscles aiding nerve innervation stability [36]. Notably, HFSC progeny secrete Sonic Hedgehog, directing the formation of this AMP-sympathetic nerve dual-component niche [36]. These evolving researches are progressively elucidating the complexities of epithelial-mesenchymal feedback loops.

Factors influencing epidermal stem cell behavior

EpSCs are regulated by a spectrum of psychological and physiological factors, including stress, mechanical forces, aging, and circadian rhythms. Understanding these influences is crucial for identifying the underlying causes of various skin disorders and developing effective treatments (Fig. 3, Table 1).

Fig. 3
figure 3

Factors influencing epidermal stem cell behavior. Psychological stress, mechanical force, chronic aging, and circadian rhythm affect various activities of EpSCs. It includes the balance between cell proliferation and differentiation, the restricted movement and abnormal migration of EpSCs from the niche, as well as the necrosis and programmed cell death of them. Consequently, processes of epithelial renewal and hair regeneration may be altered by these factors, manifesting as abnormalities in the skin like cutaneous disorders, and symptoms of skin aging such as epidermal atrophy

Table 1 Factors influencing epidermal stem cell behavior

Impact of psychological stress on epidermal stem cell functionality

Psychological stress can be broadly defined as an actual or anticipated disruption of homeostasis or an anticipated threat to well-being [37], which can be broadly divided into two types: acute and chronic. The former may activate the autonomic nervous system, mainly the sympathetic nerves, releasing noradrenaline in a burst manner [38], leading to hair greying through ectopic differentiation and depletion of MeSCs, both in mice [39] and ex vivo cultured human hair [40]. Chronic, sustained exposure to stressors, however, function primarily by activation of the hypothalamic-pituitary-adrenocortical axis and the consequently elevated circulating glucocorticoids [41]. Corticosterone in mice acts on the dermal papillae to suppress the expression of Gas6, resulting in prolonged quiescence of HFSCs and extended resting phase of HF [42]. Paradoxically, it has been reported that skin-resident regulatory T cells can respond to increased glucocorticoid signal to facilitate HFSC activation and HF regeneration through glucocorticoid receptor (GR)-transforming growth factor-β3 (TGF-β3) axis [43], which may suggest that various concentrations, sources, and exposure duration of glucocorticoid act on different cell types, so as to affect HFSCs in distinctive ways. Collectively, researches so far have identified HF as the core target of stress hormones in the epidermis, revealing the mechanism of stress-dependent hair loss and greying. Future studies may help to delineate the impact of psychological stress on other stem cell clusters in the epidermis, probably through stress-induced inflammation, which has been elucidated in the intestine [44], and interaction among EpSCs, nerve fibers, and immune cells in their niches.

The role of mechanical forces in regulating epidermal stem cell proliferation and differentiation

The epidermis, as the body's outermost barrier, is subject to a variety of mechanical forces, defined as forces resulting from contact and altering motion or rest states [45]. Common forms include compression and tensile stress, which can fluctuate with physiological and pathological processes like skin renewal and hair loss [46, 47]. Mechanical forces affecting cells originate either directly from the cytoskeleton and cell–cell adhesion in basal layers, or indirectly from the niche, impacting EpSCs [48, 49]. These forces are critical in regulating EpSC morphology and gene expression, influencing the balance between proliferation and differentiation. This regulation occurs through mechanosensory and mechanotransduction proteins within the epidermis [50, 51].

Mechanical feedback often acts as a suppressor of EpSC division [46]. Under mechanical compression, B-plexin, a transmembrane receptor, detects the force, subsequently inhibiting tissue overgrowth by repressing the transcriptional co-activator yes-associated protein 1 (YAP) [52]. THY1, integral to adherens junctions, also contributes to YAP repression [53]. Notably, in procedures like reconstructive surgery, stretching of the epidermis temporarily skews EpSCs towards renewal, mediated by activation of the Hippo and EGFR (epidermal growth factor receptor)-Ras-MAPK (Mitogen-activated protein kinase) pathways [54, 55]. Moreover, epidermal differentiation can be impeded by tension transferred to the nucleus through the linker of nucleoskeleton and cytoskeleton (LINC), which conveys mechanical forces from integrins to the cell membrane [56].

Mechanical forces originating from the differentiated epidermis and dermis indirectly influence the fate of EpSCs. It has been shown that increased contractility in the upper layers can induce intra-tissue tension, stimulating EpSC proliferation while inhibiting their differentiation and migration [57]. In conditions like aging and androgenic alopecia, hair shaft miniaturization reduces the physical niche size, leading to HFSC depletion via the Piezo1-calcium-TNF-α (tumor necrosis factor α) signaling pathway [47]. Interestingly, Piezo1 receptors of EpSCs also respond to stretching forces from adjacent dermal cell movements, enhancing regeneration in organoid culture [58]. This suggests that a single mechanosensitive receptor can produce different outcomes based on the stimulus. Overall, compression and tensile stress exert contrasting effects on EpSCs: compression typically causes depletion and suppressed division, while stretching promotes proliferation and hair regeneration. Given their safety and simplicity, the strategic application of mechanical forces holds promise for regenerative medicine research and applications.

Influence of chronic aging on epidermal stem cell dynamics

Aging, a progressive physiological process, can be divided into chronological aging influenced by intrinsic factors and extrinsic aging like photoaging [59]. EpSCs play a key role in aging, leading to visible skin changes such as thinning, hair loss, greying, and decreased dermal elasticity, which results in wrinkles [2]. The debate continues on whether a reduction in EpSC proliferative capacity contributes to skin aging [60,61,62]. Aging's effects on EpSCs can be divided into intrinsic changes affecting cell behavior directly and extrinsic changes impacting them indirectly, including altered physical niches, inflammatory cytokine release, and signaling inhibitor secretion [63].

Intrinsic alterations in epidermal stem cells due to aging

Aging has been shown to modify factors within Wnt and bone morphogenetic protein (BMP) signaling pathways, pivotal for HFSC activation and quiescence, thereby influencing HFSC functionality [64, 65]. For instance, aged HFSCs exhibit a lack of nuclear factor of activated T cells 1 (Nfatc1) deacetylation due to reduced Sirt7, impeding the transition from the telogen to anagen phase and causing delayed hair growth [66]. Aging is also marked by diminished expression of cell adhesion and extracellular matrix (ECM) genes in EpSCs, notably changes in collagen type XVII alpha 1 (COL17A1), a hemidesmosome component [67]. Both chronological aging and photoaging prompt a spread of COL17A1high cells in the interfollicular epidermis, eventually depleting EpSCs [68]. Furthermore, aging-induced hemidesmosome instability due to COL17A1 proteolysis leads to asymmetric HFSC division, migration, differentiation, and depletion [69]. Foxc1 and Nfatc1-mediated alterations in cell adhesion and ECM are implicated in EpSC migration in aged mice [70]. Overall, aging-induced changes in gene expression affect EpSC interactions with their niches, altering division and motility, contributing to epidermal atrophy, fragility, and hair thinning. This suggests COL17A1 as a potential target for reversing age-associated disorders in clinical or cosmetic applications.

The compromising effect of an aged niche on epidermal stem cells

Environmental alterations linked to aging, particularly those within the niche, exert a significant impact on the dynamics of EpSCs. Recent studies have shown that the rejuvenation of aged EpSCs is possible when paired with young dermis in transplantation assays. Conversely, young EpSCs lose regenerative capacity when combined with aged dermis, highlighting the critical role of the niche in EpSC regulation and skin aging [71]. Mechanistically, an age-associated secretory molecule in dermal fibroblasts, pentraxin 3, leads to dermal sclerotization, stimulating Piezo1 in IFESCs and causing dysregulation [72]. Furthermore, microenvironment stiffening during aging impairs HFSC function by altering the accessibility of bivalent promoters, dependent on the niche [73]. Aging also impairs dermal papilla function, diminishing HF regeneration and contributing to age-related hair loss [74]. In photoaging, a reduced level of laminin-511 at the dermal–epidermal junction is linked to a decrease in MCSP+ (Melanoma-associated chondroitin sulphate proteoglycan) Krt15+ EpSCs in humans [75].

Elevated inflammatory cytokines and secreted factors during aging are reported to affect EpSCs by regulating signaling pathways. Recent single-cell sequencing data has verified the elevation of inflammatory response gene in IFESCs during aging, including S100A9 and S100A8 [76], and global repression of cytokine signaling through Jak-STAT inhibition has a proproliferative effect directly on aged EpSCs [60], Additionally, an upregulated expression level of canonical Wnt signaling inhibitor Dkk1 and Sfrp4, as well as a downregulated expression level of BMP signaling inhibitor follistatin have been reported to retain telogen phase of hair cycle in aged mouse [77]. It follows that, through direct alterations in structure or components in niches, aging may direct the dynamic of EpSCs in an indirect way, which offers more options for clinically beneficial targets in anti-aging therapies.

The coordination of epidermal stem cell activities by circadian rhythms

The circadian system, including the central pacemaker in the suprachiasmatic nucleus and peripheral clocks like the epidermis, regulates physiological homeostasis in organisms [78, 79]. The core circadian transcription factor complex brain and muscle ARNT-like 1 (BMAL1)/circadian locomotor output cycles kaput (CLOCK), along with clock-controlled genes such as period (Per) and cryptochrome (Cry), modulate the expression of EpSC regulatory genes, influencing cell behavior [80]. BMAL1 deletion leads to an accumulation of reactive oxygen species (ROS) and continuous, unvaried cell proliferation [81]. Notably, circadian rhythms affect the rate, not the phase, of EpSC proliferation [82]. Contrarily, other studies show that BMAL1 depletion impairs proliferation and migration following injury, increases ROS levels, and delays wound healing, possibly due to ROS dual functions and variances in mouse age across studies [83]. Disruptions in circadian rhythms, such as overexpressing PER1/PER2 or deleting CRY1/CRY2, also induce premature differentiation [84], while DNA damage and repair are regulated by clock genes[85]. Intriguingly, the timing of maximum ultraviolet radiation B (UVB) exposure and the peak S-phase activity of human EpSCs in the late afternoon may contribute to higher skin cancer risks [85].

Circadian rhythms orchestrate the activities of EpSCs through innervation from the suprachiasmatic nucleus and peripheral tissue signals. Recent findings reveal that light exposure prompts the suprachiasmatic nucleus to activate HFSCs via sympathetic nerves and downstream hedgehog signaling [86]. Additionally, peripheral circadian clocks can respond to light and sustain normal diurnal rhythms independently of the central pacemaker, possibly via photopigments in epidermal keratinocytes [87]. Notably, clock genes exhibit uniform expression in HF progenitor cells in both humans and mice but show a patchy pattern in bulge HFSCs, suggesting that HFSCs may be at different circadian phases, leading to varied responses to circadian cues [88].

Regulatory mechanisms of epidermal stem cell activities

EpSCs are central to skin homeostasis, where their proliferation, differentiation, and decision between activation and quiescence are intricately regulated. Critical to this process is the delicate equilibrium between cell number increase and fate determination, influenced by classical signaling pathways and various key molecules (Fig. 4, Table 2).

Fig. 4
figure 4

Regulatory mechanisms of epidermal stem cell activities. The delicate equilibrium of EpSC between cell number increase and fate determination is majorly influenced by four classical signaling pathways and various key molecules. The horizontal and vertical axes represent “proliferation” and “differentiation” respectively. To elaborate, the red arrows of which represent “promote” and the suppression symbols in black represent “inhibit”. Signaling pathways and molecules positioned in different quadrants indicate their roles in promoting or inhibiting EpSC differentiation and proliferation. Specifically, the Wnt/β-Catenin signaling pathway is responsible for promoting both cell proliferation and differentiation. The primary role of the Sonic hedgehog (Shh) signaling pathway is to enhance cell growth, with only a modest effect on differentiation. As for the Notch and BMP pathways, they foster the differentiation of EpSCs while concurrently suppressing their proliferation. Molecules that steer the activities of EpSCs by interacting with the four signaling pathways are also illustrated in the figure

Table 2 Regulators compromising epidermal stem cell activities

Impact of classical signaling pathways on the fate of epidermal stem cells

The activities of EpSCs are naturally controlled by a series of signaling pathways under physiological conditions, and the alteration of which usually leads to pathological states [89, 90]. Among them, the Wnt/β-catenin pathway, Sonic hedgehog (Shh) pathway, Notch pathway, and BMP pathway are four major regulating mechanisms that determine the fate of EpSCs collectively [8].

The Wnt/β-catenin pathway

Wnt proteins, belonging to a large family of extracellular signals, are crucial for developmental processes. They engage Frizzled receptors on cell surfaces, triggering reactions that stabilize β-catenin in the cytoplasm, thus initiating transcription [91]. The Wnt/β-catenin pathway is key in driving the transition of HFs from the resting phase (telogen) to the growth phase (anagen), facilitating both proliferation and differentiation [92]. Recent studies have also identified the receptor tyrosine kinase-like orphan receptor 2 (ROR2) as an additional functional receptor in Wnt signaling, expanding the understanding of β-catenin-dependent pathways [90].

The Sonic hedgehog pathway

Shh, an extracellular protein like Wnt, is vital for animal development [93]. It functions by binding to its receptor Patched (Ptch), thereby liberating Smoothened (Smo) from Ptch inhibition. This activation relieves the suppressor of fused homologue (Sufu) inhibition on the transcription factor Gli, leading to the activation of target genes [94]. Shh signaling plays a crucial role in stimulating the proliferation of both quiescent EpSCs and TACs [8]. Dysregulation of Shh signaling in EpSCs is implicated in severe skin disorders, including basal cell carcinoma and hair loss [95, 96].

The Notch pathway

Notch is a transmembrane receptor that modulates various cell fate decisions across species [97]. Functional Notch receptor binds to a ligand on a neighboring cell, then it is successively cleaved by a disintegrin and metalloprotease and γ-secretase into the active intracellular domain, which translocate to the nucleus and induces transcription with coactivators [97]. The Notch pathway is highly related to the differentiation of EpSCs, while its impacts on cell proliferation are seemingly distinct in different parts of the epidermis [8, 98]. Moreover, its function in regulating metabolism in HFSCs has been reported [99].

The bone morphogenetic protein pathway

BMPs, part of the TGF-β superfamily, are secreted proteins often modulated by inhibitors like Noggin [100]. Upon binding to specific receptors, BMPs trigger the phosphorylation and activation of R-Smad in the cytoplasm, which then associates with co-Smad partner Smad 4, initiating transcription [100]. BMP signaling primarily maintains EpSCs in a quiescent state, inhibiting self-renewal and differentiation [101, 102]. However, it is also pivotal in guiding activated stem cells towards specific lineages [103]. Notably, growth differentiation factor 5 (GDF-5), also known as BMP-14, has been shown to enhance mouse EpSC proliferation through the Foxg1-cyclinD1 axis [104].

Emerging insights into cell metabolism in regulating epidermal stem cell states

Recent findings highlight cell metabolism as a key regulator of EpSCs, influencing their state and functions including energy production, signaling, and gene expression [105]. HFSCs in hypoxic niches primarily rely on anaerobic glycolysis, minimizing ROS damage [106]. During differentiation, they switch to oxidative phosphorylation, using glutamine as an additional carbon source to optimize ATP production [107]. Inhibiting glutaminase via mTORC2-Akt signaling has been shown to revert progenitor cells to a stem cell state post-anagen phase [108]. Additionally, glutamine enhances skin and HF regeneration under bacterially induced hypoxia via interleukin (IL)-1β-MyD88 signaling, demonstrating its complex role in metabolism [109]. Furthermore, various lipids impact EpSC proliferation and differentiation. For instance, free fatty acids in high-fat diets promote ROS accumulation and short-term epidermal differentiation, but long-term can deplete HFSCs by inhibiting the Shh pathway through NFκB (nuclear factor kappa-B) and IL-1R (type I IL-1 receptor) signaling [95]. Ceramides and glucosylceramides also induce keratinocyte differentiation, though the specific mechanisms remain unexplored [110]. The dysregulation of differentiation is closely linked to oncogenesis, with abnormal EpSC behavior being a potential root of malignancy [111]. Recent studies suggest that limiting extracellular serine encourages EpSCs to stimulate de novo serine synthesis, activating differentiation programs and potentially advancing malignancy, offering new insights into targeting oncogenic stem cells [112]. Apart from metabolites derived from the three primary nutrients, recent research indicates that all-trans retinoic acid (atRA), a metabolite of vitamin A, plays a crucial role in resolving lineage plasticity following injury. Specifically, atRA promotes the identity of HFSCs while simultaneously suppressing the fate of EpSCs at the chromatin level [113].

Other key molecules that govern the identity and regenerative capacity of epidermal stem cells

In addition to metabolites, the fate determination, clonal proliferation, and even apoptosis of EpSCs are influenced by a diverse array of cell-intrinsic and extrinsic molecules. These include transcription factors, epigenetic regulators, membrane receptors, and cytokines [114]. Many of these molecules act by modulating the four established signaling pathways, while others exert direct and independent effects. Recent studies have demonstrated that upon injury, EpSCs express and respond to IL-24 through both autocrine and paracrine mechanisms to proliferate. This process also stimulates the proliferation of dermal endothelial cells and fibroblasts [115]. Remarkably, this molecular cascade bears similarities to immune responses triggered by microbes, B cells, or T cells, highlighting its distinct yet functionally analogous nature. These newly identified regulators have unveiled novel functionalities and potent capabilities. However, further research is necessary to uncover the specific triggers that initiate changes in these molecules during the anagen or catagen phases, and to elucidate the intricate crosstalk among these key modulators.

Epidermal stem cells in immune-related skin diseases

EpSCs play a crucial role in re-epithelialization during wound healing, a key aspect of skin homeostasis. Dysfunctional EpSCs are also implicated in various skin disorders including psoriasis and vitiligo. Moreover, the concept of 'inflammatory memory' in EpSCs has emerged as a significant factor in both normal and pathological skin processes [116] (Fig. 5).

Fig. 5
figure 5

Epidermal stem cells in wound healing and immune-related skin diseases. a Wound Healing: epidermal stem cells (EpSCs) from various hair follicle (HF) regions, including the infundibulum, isthmus, bulge, and sebaceous glands, aid in repairing the interfollicular epidermis (IFE). Under stress, HFSCs transition towards IFESCs, expressing markers of both populations during migration and retaining epigenetic changes for a rapid response to future injuries. Additionally, stem cells and progenitors in the IFE and cells from the sweat gland duct contribute to wound healing. b Hair Loss: the dysfunction or loss of HFSCs, often due to inflammation in the bulge or dermal papilla, oxidative stress, and disrupted gene regulation, is a key factor in hair loss. c Melanocyte Homeostasis: EpSCs, co-existing with melanocyte stem cells in the bulge, are crucial for melanocyte health. Dysfunctional EpSCs can lead to melanocyte damage and the formation of white patches. d Psoriasis and EpSCs: the 'inflammatory memory' of EpSCs, characterized by enhanced transcriptional responses to inflammation, contributes to the pathology of psoriasis. e Cutaneous lupus erythematosus: inflammation-related damage to HFSCs, particularly in the bulge area, may cause permanent hair loss in this condition. f Hidradenitis Suppurativa: cells from the outer root sheath in affected patients show an increased number of proliferating progenitors and fewer quiescent stem cells. This imbalance contributes to inflammation through the promotion of type I interferon synthesis

Wound healing

The wound healing process in the skin involves a coordinated effort of various tissue components, with EpSCs playing a central role. This process encompasses four stages: hemostasis, inflammation, proliferation, and remodeling [4]. Hemostasis involves forming a coagulation and fibrin clot that supports cellular migration and immune cell recruitment [117]. During the second stage, immune cells clear pathogens and release cytokines to activate skin stem/progenitor cells like EpSCs, enhancing cell proliferation, differentiation, and shaping the regenerative microenvironment [118]. For instance, IL-1α is known to accelerate the regeneration of Lgr5+ HFSCs and initiate HF formation, though the origin of regenerated HFSCs is unclear [119]. EpSCs, in turn, secrete Ccl2 to influence macrophage behavior, facilitating tissue repair via keratinocyte-macrophage interactions [120]. Subsequently, the proliferation stage requires collaboration between keratinocytes and fibroblasts for re-epithelialization and dermal repair [121]. Finally, in the remodeling stage, excess cells are removed, and the ECM in the wound bed is restructured [117].

During wound healing, EpSCs are crucial in the proliferation phase, contributing to re-epithelialization via migration, population dynamics, and plasticity [67, 122]. Intriguingly, EpSC migration and proliferation occur in distinct zones: rapid migration without proliferation at the wound's leading edge, and concentrated proliferation with minimal migration in a distant zone [4]. Migration is regulated by mechanisms such as the Piezo1 ion channel, which when activated at the cell's rear, induces cell retraction and slows migration, thus impacting wound healing [123]. Further research is required to fully understand these dynamics. Additionally, EpSCs adapt by increasing symmetric cell division and decreasing differentiation to replenish lost cells [124].

EpSCs exhibit remarkable plasticity, enabling them to rapidly alter their fate for effective tissue regeneration. Originally, HFSCs were believed to contribute to wound repair 2–3 days post-injury [125]. However, recent evidence indicates that HFSCs commence fate determination within 24 h post-injury, even before leaving the bulge area [126]. Under stress, migrating HFSCs enter a state of 'lineage infidelity' near wounds, expressing markers of both IFESCs and HFSCs, driven by factors like ETS2 [127, 128]. They also retain an epigenetic memory of the wound, which enhances their responsiveness to subsequent injuries [129]. Furthermore, single-cell RNA sequencing has revealed that IFESCs undergo a progression from a quiescent state in the proliferation zone to a more differentiated state in the migration zone, demonstrating a unique aspect of stem cell plasticity [14]. In more severe wounds, EpSCs near blood vessels can differentiate into vascular endothelial cells, aiding angiogenesis [130]. While this plasticity is promising for regenerative medicine, its potential role in tumorigenesis warrants careful consideration [125].

Hair loss

Hair cycle dysregulation often leads to hair loss, primarily caused by impaired functions or irreversible damage to HFSCs, such as apoptosis, depletion, and abnormal differentiation. These issues can disrupt the normal hair growth cycle, leading to conditions like permanent scarring alopecia [131]. Contributing factors include hormonal imbalances, autoimmune and nutritional disorders, genetic predisposition, radiation, chemotherapy, and aging [131, 132]. Scarring alopecia, for example, is marked by excessive autoimmune responses in the HF bulge and infiltration of Th1 cytotoxic T cells [133]. Early inflammation in the dermal papilla or other HF components affecting HFSC activities can lead to reversible hair loss forms like alopecia areata and androgenetic alopecia [134]. Although the exact cellular and molecular mechanisms of hair loss remain elusive, ongoing research is exploring various triggers and pathways potentially impacting EpSC proliferation and differentiation, as discussed above (Table 2).

Psoriasis

Psoriasis, a chronic inflammatory skin condition, affects over 60 million people globally and is closely linked to the behavior of EpSCs [135]. Disruptions in the balance between EpSC proliferation and differentiation, often regulated by key factors like p63, are associated with psoriatic conditions. Depletion of the primary p63 isoform ΔNp63 leads to a less proliferative keratinocyte population, triggering an inflammatory, psoriasis-like state, while its overexpression may suppress this condition [136]. Similarly, the knockout of ΔNp63 effectors c-JUN and JUNB results in psoriasis-like symptoms, with mutant HFSCs compensating by stimulating the proliferation of neighboring cells [137]. The role of EpSC subsets in psoriasis severity and types is also notable. For instance, deleting the Wnt receptor Evi/Wls in Krt14-expressing IFESCs and progenitors induces a psoriasiform dermatitis-like phenotype in mice [138]. Additionally, imiquimod treatment in HFs can trigger premature hair cycle entry through Wnt-independent mechanisms [139].

Moreover, the 'inflammatory memory' of EpSCs is a significant factor in psoriasis pathogenesis. This memory, characterized by sustained chromosomal accessibility for rapid response to secondary challenges, is influenced by factors like absent in melanoma 2 (AIM2) and transcription factors FOS and JUN, which collaborate with stimulus-specific STAT3 [140, 141]. This heightened sensitivity of EpSCs to tissue damage might partly explain the excessive immune reactions observed in psoriasis and other recurrent inflammatory skin diseases [142]. Targeting the inflammation-rewired EpSCs could be a potential therapeutic strategy for these conditions.

Other inflammatory skin diseases

EpSCs dysfunction is increasingly linked to various inflammatory skin conditions, including cutaneous lupus erythematosus, hidradenitis suppurativa, and vitiligo. In cutaneous lupus erythematosus, inflammation in the HF bulge area can lead to permanent hair loss, akin to some forms of alopecia [143]. Research indicates that in these cases, HFSCs are either repurposed for repair or damaged by secondary inflammatory responses [144]. Conversely, in hidradenitis suppurativa, an inflammatory disease affecting the pilosebaceous-apocrine unit, HFSCs have been found to produce type I interferon (IFN) [145, 146]. This condition is also marked by increased HFSC proliferation and differentiation. The underlying mechanism involves replication stress in HFSCs due to stalled replication forks, activating the ATR/CHK1 pathway, which in turn stimulates IFN synthesis via the IFI16/STING pathway [146].

Vitiligo, an autoimmune skin disorder characterized by depigmentation, involves complex interactions between keratinocytes and melanocytes. Keratinocytes play a crucial role in maintaining melanocyte homeostasis and melanogenesis, and in modulating immune responses [147]. They support melanocyte function by providing essential growth factors and regulate melanocyte activity through metalloproteinases and basement membrane remodeling [148]. However, in vitiligo, dysfunctional keratinocytes can exacerbate autoimmune reactions by presenting melanocyte antigens and releasing pro-inflammatory cytokines and chemokines in affected areas [149]. The presence of abnormal keratinocytes and melanocytes in vitiligo lesions further underscores the pathogenic role of epidermal cells [147]. Research also suggests the potential of leveraging EpSCs in vitiligo treatment through the melanocyte-keratinocyte system.

Clinical applications of epidermal stem cells

Since the initial cultivation of human skin stem cells in vitro and their successful engraftment in murine models [150,151,152], EpSCs have become a focal point in research and clinical applications for wound management and other skin disorders, thanks to their accessibility and regenerative properties [153]. Additionally, the secreted factors and extracellular vesicles from EpSCs show great potential in novel therapeutic approaches for wound healing and hair regeneration [154].

Epidermal stem cells in wound repair and alopecia treatment

In recent decades, cultured epidermal autografts enriched with EpSCs and HF transplantation have advanced wound repair and alopecia treatments [153, 155]. However, challenges persist: cultured epidermal autografts currently lack epidermal appendages like sweat glands, and EpSC proliferation is time-consuming for patients in urgent need of treatment [153, 156]; in alopecia treatment, the availability of transplantable HFs is limited, and transplanted HFs risk dysfunction due to unchanged microenvironment [132]. Innovations are addressing these challenges. Tissue-engineering approaches, such as skin and HF organoids, offer potential solutions, especially in inducing de novo regeneration of HFs and sebaceous glands using dermal papilla cells and skin-derived precursors [157, 158]. Following the discovery that skin organoids can be developed from human pluripotent stem cells [159], recent researches using human induced pluripotent stem cells (iPSCs) have successfully developed skin organoids with sweat glands and Merkel cells [160]. The optimized protocol exhibited greater clinical application value while tackling the ethical concerns surrounding fetal tissue-derived stem cells. As for the delay in treatment, Rho-associated kinase inhibitors have shown promise in expediting EpSC proliferation for cultured epidermal autografts [161]. Recent advancements include biomimetic ECMs that regulate HFSC fate, eliminating the need for feeder cells [162]. Overall, further research is necessary to develop vascularized skin organoids that incorporate immune cells. This advancement holds potential to address challenges such as large non-healing skin wounds in hypoxic conditions, deficiencies in micronutrients like those seen in diabetic ulcers, and the dysregulated immune microenvironment in alopecia patients. To achieve clinical applicability, it is essential to establish standardized cell sources, consistent differentiation protocols, and robust validation assays for the functional capabilities of EpSCs.

3D bioprinting, encompassing techniques like inkjet, laser-assisted, extrusion, and scaffold-free spheroid-based bioprinting, is revolutionizing bioengineering by creating precise skin organoids with functional appendages [163,164,165]. These bioinks, combining EpSCs, mesenchymal cells, and biomaterials, enhance the viability and differentiation of EpSCs. Innovations in biomaterial scaffolds, including nanoscale biomimetic ECM, molecular hydrogen, and platelet-rich plasma, are improving microenvironment for EpSC regeneration and angiogenesis [162, 166, 167]. Safety and effectiveness of EpSC therapies are being validated through preclinical studies and deep learning-based cell tracking technologies [168, 169].

Paracrine factors from EpSCs, contributing significantly to their regenerative potential, may be utilized independently as cell-free therapy that promote angiogenesis and epithelialization in skin defect treatments [155, 170]. EpSC-derived extracellular vesicles, enhancing both EpSC and dermal cell proliferation, are emerging as promising therapies for wound repair and tissue regeneration [154, 171, 172].

Transgenic epidermal stem cell therapy for epidermolysis bullosa treatment

Epidermolysis bullosa (EB) is a severe skin disorder arising from genetic defects in the dermo-epidermal junction [173]. Classified mainly into simplex, junctional, dystrophic, and mixed types based on blister formation, EB requires different treatments tailored to specific molecular and genetic abnormalities [173, 174]. Transgenic EpSC transplantation has shown remarkable results in EB treatment. In 2006, a junctional EB patient received transgenic EpSCs, demonstrating stable and durable skin improvement over 16 years [175, 176]. A notable 2017 case involved replacing 80% of a patient's skin with transgenic epidermis, resulting in a fully functional skin barrier maintained over six years [177, 178]. Laminin subunit beta 3 (LAMB3)-deficiency in junctional EB, leading to reduced YAP activity in EpSCs and subsequent cell depletion, presents a new gene therapy target [179]. In recessive dystrophic EB, autologous cultured transgenic EpSCs have restored collagen VII expression in patients, with over 70% of treated wounds healed and maintaining collagen VII expression after two years [180, 181]. However, gene modification techniques, such as integrating vectors carrying functional genes, have proven effective primarily for recessively inherited EB. For EB simplex, which is dominantly inherited, there is a critical need for integrated ex vivo cell and genetic therapies to precisely correct the mutated allele. This approach holds promise for advancing treatment options in these cases.

Emerging gene-editing technologies like CRISPR/Cas9 offer potential for correcting both dominant and recessive mutations [182, 183]. Recent studies have demonstrated targeted cleavage of the mutant allele in EB simplex, restoring a normal cellular phenotype without affecting the normal functions of EpSCs, which represents a significant advancement in precision medicine for treating EB simplex [184]. Although genetically edited EpSCs are mainly in the preclinical stage, their full development holds significant promise for EB treatment. To this end, future studies are expected to optimize gene-editing tools, such as prime editing, base editing, and Cpf1/Cas12-based editing, to attain greatest on-target efficiency and minimal off-target effects. Notably, genotoxicity and genomic instabilities are two serious issues that requires careful prevention and thorough verification.

Cultured epidermal stem cells with melanocytes for vitiligo treatment

EpSCs are integral to vitiligo treatments, encompassing tissue and cellular grafting methods [185]. The melanocyte-keratinocyte co-cultured system, initially used in autologous cultured epidermal grafting for stable vitiligo, achieved significant permanent repigmentation [186]. Subsequently, non-cultured epidermal cell grafts emerged as a rapid and efficient approach for treating extensive vitiligo [187]. The autologous melanocyte-keratinocyte suspension, despite requiring larger donor biopsies, is favored for its convenience and cost-effectiveness [188].

Recent advancements show that epidermal cell suspensions combined with follicular cells, rich in MeSCs, fibroblast growth factor, and stem cell factor, lead to superior and faster repigmentation [189]. This effectiveness is likely due to the diverse types of HFSCs and MeSCs in the bulge area. The dynamic interplay between various EpSC populations and melanocytes presents significant clinical potential, suggesting that specific EpSC clusters might enhance melanocyte proliferation and improve outcomes to meet growing clinical demands.

Methods for studying epidermal stem cells

The distinctive biological properties of EpSCs, such as unique biomarkers, self-renewal, differentiation capabilities, slow-cycling nature, and adhesion traits, have paved the way for diverse research methodologies. Advances in research techniques have further deepened our understanding of EpSC biology, origins, and clinical applications (Fig. 6).

Fig. 6
figure 6

Methods for studying epidermal stem cells. EpSCs can be obtained from neonatal foreskin samples and mouse skin, enriched after 20-min incubation on collagen IV, and expanded in culture media with 3T3 fibroblasts as feeder cells. iPSCs are also optional sources of EpSCs. Specific research methods corresponding to the slow cycling nature and regenerative ability of EpSCs include label retention and skin regeneration assays. Live imaging, machine learning, and single cell RNA-sequencing are novel techniques currently in use. All the figure created using BioRender (https://biorender.com/)

Isolation, enrichment, and expansion of epidermal stem cells in vitro

Primary human epidermal keratinocytes, often sourced from neonatal foreskin, and basal layers of mouse skin are common reservoirs for EpSCs [190, 191]. EpSCs can also be derived from iPSCs, expressing CD200 and ITGA6, capable of forming all HF lineages and reconstituting the interfollicular epidermis [192]. Explant culture and enzymatic digestion are standard isolation methods, with recent studies favoring hyaluronidase and collagenase I over trypsin for higher cell viability and stemness [193]. Fluorescence-activated cell sorting is used for cell selection, identifying EpSCs through markers like high integrin β1 and Krt19/5/14 expression [194, 195]. A simple and special method for EpSC enrichment involves incubation on collagen IV, leveraging fast adhesion through integrin β1 [194].

Culture media advancements have enabled efficient and quality in vitro EpSC expansion [165]. The Rheinwald and Green method, using co-culture systems with feeder cells like 3T3-J2 and mitomycin-C, is prevalent [152]. Serum-free media with lower immunogenicity have also been developed for more refined culture conditions [196]. Except for culture media, multiple devices have popped up to optimize culture systems [197]. Finally, three-dimensional culture systems to mimic the real expansion condition of EpSCs have been developed at lightning speed, which has been discussed in the previous section.

Label retention and transplantation of epidermal stem cells in vivo

Utilizing their slow-cycling nature, EpSCs are able to be studied through label retention techniques. Radioactively-labeled nucleotides like 5-bromo-2’-deoxyuridine (BrdU) help identify EpSCs in vivo due to their infrequent division [198, 199]. The use of GFP-tagged histone H2B (H2B-GFP) in a tet-controllable manner enhances specificity in chromatin transition studies [200, 201]. Skin regeneration assays test EpSC regenerative abilities and are complemented by transplanting cultured EpSCs or organoids onto mice for more comprehensive functional assessment [58, 159].

Novel techniques in epidermal stem cell research

Advanced tools like live imaging, machine learning, and single-cell RNA-sequencing have transformed EpSC research. Genetic lineage tracing with the Cre-loxP system and intravital microscopy can capture dynamics of EpSCs via live imaging [202, 203]. Moreover, machine learning facilitates in-depth analysis of EpSC behavior and interactions with high efficiency [168]. Single-cell RNA-sequencing offers insights into lineage hierarchies, signaling networks, cell state predictions, and stem cell quantification in cell populations [15, 25]. These methods, coupled with the evolution of artificial intelligence and bioinformatics, are propelling significant breakthroughs in EpSC research.

Conclusions

This review provides an updated perspective on EpSCs in skin homeostasis, potential factors influencing their behavior, their involvement in clinical perspectives, as well as regulatory mechanisms and methods used for studying them. However, limitations remain in translating findings to clinical applications due to differences between species, and in the widespread adoption of EpSC-based treatments because of high costs and associated risks. With the aid of rapidly evolving techniques, it is likely that we will gain a more comprehensive understanding EpSCs in the near future.

Availability of data and materials

Not applicable.

Abbreviations

AIM2:

Absent in melanoma 2

atRA:

All-trans retinoic acid

BMAL1:

Brain and muscle ARNT-like 1

BMP:

Bone morphogenetic protein

BrdU:

5-Bromo-2’-deoxyuridine

CLOCK:

Circadian locomotor output cycles kaput

COL17A1:

Collagen type XVII alpha 1

Cry:

Cryptochrome

EB:

Epidermolysis bullosa

ECM:

Extracellular matrix

EGFR:

Epidermal growth factor receptor

EpSC:

Epidermal stem cell

GDF-5:

Growth differentiation factor 5

GR:

Glucocorticoid receptor

H2B-GFP:

GFP-tagged histone H2B

HF:

Hair follicle

HFSC:

Hair follicle stem cell

IFESC:

Interfollicular epidermal stem cell

IFN:

Interferon

IL:

Interleukin

IL-1R:

IL-1 receptor

iPSCs:

Induced Pluripotent stem cells

Krt:

Keratin

LAMB3:

Laminin subunit beta 3

LINC:

Linker of nucleoskeleton and cytoskeleton

MAPK:

Mitogen-activated protein kinase

MCSP:

Melanoma-associated chondroitin sulfate proteoglycan

MeSC:

Melanocyte stem cell

Nfatc1:

Nuclear factor of activated T cells 1

NFκB:

Nuclear factor kappa-B

Per:

Period

Ptch:

Patched

ROR2:

Receptor tyrosine kinase-like orphan receptor 2

ROS:

Reactive oxygen species

Shh:

Sonic hedgehog

Smo:

Smoothened

Sufu:

Suppressor of fused homologue

TAC:

Transit-amplifying cell

TGF-β3:

Transforming growth factor-β3

TNF-α:

Tumor necrosis factor α

UVB:

Ultraviolet radiation B

YAP:

Yes-associated protein

References

  1. Blanpain C, Fuchs E. Epidermal homeostasis: a balancing act of stem cells in the skin. Nat Rev Mol Cell Biol. 2009;10(3):207–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Hsu YC, Li L, Fuchs E. Emerging interactions between skin stem cells and their niches. Nat Med. 2014;20(8):847–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Solanas G, Benitah SA. Regenerating the skin: a task for the heterogeneous stem cell pool and surrounding niche. Nat Rev Mol Cell Biol. 2013;14(11):737–48.

    Article  CAS  PubMed  Google Scholar 

  4. Dekoninck S, Blanpain C. Stem cell dynamics, migration and plasticity during wound healing. Nat Cell Biol. 2019;21(1):18–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Watt FM. Mammalian skin cell biology: at the interface between laboratory and clinic. Science. 2014;346(6212):937–40.

    Article  CAS  PubMed  Google Scholar 

  6. Gonzales KAU, Fuchs E. Skin and its regenerative powers: an alliance between stem cells and their niche. Dev Cell. 2017;43(4):387–401.

    Article  CAS  PubMed  Google Scholar 

  7. Stanley JR. Synergy of understanding dermatologic disease and epidermal biology. J Clin Invest. 2012;122(2):436–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Blanpain C, Fuchs E. Epidermal stem cells of the skin. Annu Rev Cell Dev Biol. 2006;22:339–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Pastushenko I, Prieto-Torres L, Gilaberte Y, Blanpain C. Skin stem cells: at the frontier between the laboratory and clinical practice. Part 1: epidermal stem cells. Actas Dermosifiliogr. 2015;106(9):725–32.

    Article  CAS  PubMed  Google Scholar 

  10. Lechler T, Fuchs E. Asymmetric cell divisions promote stratification and differentiation of mammalian skin. Nature. 2005;437(7056):275–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Damen M, Wirtz L, Soroka E, Khatif H, Kukat C, Simons BD, et al. High proliferation and delamination during skin epidermal stratification. Nat Commun. 2021;12(1):3227.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Clayton E, Doupé DP, Klein AM, Winton DJ, Simons BD, Jones PH. A single type of progenitor cell maintains normal epidermis. Nature. 2007;446(7132):185–9.

    Article  CAS  PubMed  Google Scholar 

  13. Mascré G, Dekoninck S, Drogat B, Youssef KK, Broheé S, Sotiropoulou PA, et al. Distinct contribution of stem and progenitor cells to epidermal maintenance. Nature. 2012;489(7415):257–62.

    Article  PubMed  Google Scholar 

  14. Haensel D, Jin S, Sun P, Cinco R, Dragan M, Nguyen Q, et al. Defining epidermal basal cell states during skin homeostasis and wound healing using single-cell transcriptomics. Cell Rep. 2020;30(11):3932-3947.e6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Wang S, Drummond ML, Guerrero-Juarez CF, Tarapore E, MacLean AL, Stabell AR, et al. Single cell transcriptomics of human epidermis identifies basal stem cell transition states. Nat Commun. 2020;11(1):4239.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Singh R. Basal cells in the epidermis and epidermal differentiation. Stem Cell Rev Rep. 2022;18(6):1883–91.

    Article  PubMed  Google Scholar 

  17. Roy E, Neufeld Z, Cerone L, Wong HY, Hodgson S, Livet J, et al. Bimodal behaviour of interfollicular epidermal progenitors regulated by hair follicle position and cycling. Embo j. 2016;35(24):2658–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Cockburn K, Annusver K, Gonzalez DG, Ganesan S, May DP, Mesa KR, et al. Gradual differentiation uncoupled from cell cycle exit generates heterogeneity in the epidermal stem cell layer. Nat Cell Biol. 2022;24(12):1692–700.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Chen CL, Huang WY, Wang EHC, Tai KY, Lin SJ. Functional complexity of hair follicle stem cell niche and therapeutic targeting of niche dysfunction for hair regeneration. J Biomed Sci. 2020;27(1):43.

    Article  PubMed  PubMed Central  Google Scholar 

  20. Peterson A, Nair LS. Hair follicle stem cells for tissue regeneration. Tissue Eng Part B Rev. 2022;28(4):695–706.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Fuchs E. Scratching the surface of skin development. Nature. 2007;445(7130):834–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Blanpain C, Lowry WE, Geoghegan A, Polak L, Fuchs E. Self-renewal, multipotency, and the existence of two cell populations within an epithelial stem cell niche. Cell. 2004;118(5):635–48.

    Article  CAS  PubMed  Google Scholar 

  23. Snippert HJ, Haegebarth A, Kasper M, Jaks V, van Es JH, Barker N, et al. Lgr6 marks stem cells in the hair follicle that generate all cell lineages of the skin. Science. 2010;327(5971):1385–9.

    Article  CAS  PubMed  Google Scholar 

  24. Morita R, Sanzen N, Sasaki H, Hayashi T, Umeda M, Yoshimura M, et al. Tracing the origin of hair follicle stem cells. Nature. 2021;594(7864):547–52.

    Article  CAS  PubMed  Google Scholar 

  25. Wu S, Yu Y, Liu C, Zhang X, Zhu P, Peng Y, et al. Single-cell transcriptomics reveals lineage trajectory of human scalp hair follicle and informs mechanisms of hair graying. Cell Discov. 2022;8(1):49.

    Article  PubMed  PubMed Central  Google Scholar 

  26. Steingrímsson E, Copeland NG, Jenkins NA. Melanocyte stem cell maintenance and hair graying. Cell. 2005;121(1):9–12.

    Article  PubMed  Google Scholar 

  27. Sun Q, Lee W, Hu H, Ogawa T, De Leon S, Katehis I, et al. Dedifferentiation maintains melanocyte stem cells in a dynamic niche. Nature. 2023;616(7958):774–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Wang X, Ramos R, Phan AQ, Yamaga K, Flesher JL, Jiang S, et al. Signalling by senescent melanocytes hyperactivates hair growth. Nature. 2023;618(7966):808–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Chen R, Zhu Z, Ji S, Geng Z, Hou Q, Sun X, et al. Sweat gland regeneration: current strategies and future opportunities. Biomaterials. 2020;255:120201.

    Article  CAS  PubMed  Google Scholar 

  30. Lu CP, Polak L, Rocha AS, Pasolli HA, Chen SC, Sharma N, et al. Identification of stem cell populations in sweat glands and ducts reveals roles in homeostasis and wound repair. Cell. 2012;150(1):136–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Ankawa R, Goldberger N, Yosefzon Y, Koren E, Yusupova M, Rosner D, et al. Apoptotic cells represent a dynamic stem cell niche governing proliferation and tissue regeneration. Dev Cell. 2021;56(13):1900-1916.e5.

    Article  CAS  PubMed  Google Scholar 

  32. Fujiwara H, Ferreira M, Donati G, Marciano DK, Linton JM, Sato Y, et al. The basement membrane of hair follicle stem cells is a muscle cell niche. Cell. 2011;144(4):577–89.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Fuchs E, Blau HM. Tissue stem cells: architects of their niches. Cell Stem Cell. 2020;27(4):532–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Heitman N, Sennett R, Mok KW, Saxena N, Srivastava D, Martino P, et al. Dermal sheath contraction powers stem cell niche relocation during hair cycle regression. Science. 2020;367(6474):161–6.

    Article  CAS  PubMed  Google Scholar 

  35. Martino P, Sunkara R, Heitman N, Rangl M, Brown A, Saxena N, et al. Progenitor-derived endothelin controls dermal sheath contraction for hair follicle regression. Nat Cell Biol. 2023;25(2):222–34.

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Shwartz Y, Gonzalez-Celeiro M, Chen CL, Pasolli HA, Sheu SH, Fan SM, et al. Cell types promoting goosebumps form a niche to regulate hair follicle stem cells. Cell. 2020;182(3):578-593.e19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Ulrich-Lai YM, Herman JP. Neural regulation of endocrine and autonomic stress responses. Nat Rev Neurosci. 2009;10(6):397–409.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Biltz RG, Sawicki CM, Sheridan JF, Godbout JP. The neuroimmunology of social-stress-induced sensitization. Nat Immunol. 2022;23(11):1527–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Zhang B, Ma S, Rachmin I, He M, Baral P, Choi S, et al. Hyperactivation of sympathetic nerves drives depletion of melanocyte stem cells. Nature. 2020;577(7792):676–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Rachmin I, Lee JH, Zhang B, Sefton J, Jung I, Lee YI, et al. Stress-associated ectopic differentiation of melanocyte stem cells and ORS amelanotic melanocytes in an ex vivo human hair follicle model. Exp Dermatol. 2021;30(4):578–87.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Goldschen L, Ellrodt J, Amonoo HL, Feldman CH, Case SM, Koenen KC, et al. The link between post-traumatic stress disorder and systemic lupus erythematosus. Brain Behav Immun. 2023;108:292–301.

    Article  CAS  PubMed  Google Scholar 

  42. Choi S, Zhang B, Ma S, Gonzalez-Celeiro M, Stein D, Jin X, et al. Corticosterone inhibits GAS6 to govern hair follicle stem-cell quiescence. Nature. 2021;592(7854):428–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Liu Z, Hu X, Liang Y, Yu J, Li H, Shokhirev MN, et al. Glucocorticoid signaling and regulatory T cells cooperate to maintain the hair-follicle stem-cell niche. Nat Immunol. 2022;23(7):1086–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Schneider KM, Blank N, Alvarez Y, Thum K, Lundgren P, Litichevskiy L, et al. The enteric nervous system relays psychological stress to intestinal inflammation. Cell. 2023;186(13):2823-2838.e20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Dupont S, Wickström SA. Mechanical regulation of chromatin and transcription. Nat Rev Genet. 2022;23(10):624–43.

    Article  CAS  PubMed  Google Scholar 

  46. Vasioukhin V, Bauer C, Degenstein L, Wise B, Fuchs E. Hyperproliferation and defects in epithelial polarity upon conditional ablation of alpha-catenin in skin. Cell. 2001;104(4):605–17.

    Article  CAS  PubMed  Google Scholar 

  47. Xie Y, Chen D, Jiang K, Song L, Qian N, Du Y, et al. Hair shaft miniaturization causes stem cell depletion through mechanosensory signals mediated by a Piezo1-calcium-TNF-α axis. Cell Stem Cell. 2022;29(1):70-85.e6.

    Article  CAS  PubMed  Google Scholar 

  48. Iskratsch T, Wolfenson H, Sheetz MP. Appreciating force and shape—the rise of mechanotransduction in cell biology. Nat Rev Mol Cell Biol. 2014;15(12):825–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Lei M, Yang L, Chuong CM. Getting to the core of the dermal papilla. J Invest Dermatol. 2017;137(11):2250–3.

    Article  CAS  PubMed  Google Scholar 

  50. Kirby TJ, Lammerding J. Emerging views of the nucleus as a cellular mechanosensor. Nat Cell Biol. 2018;20(4):373–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Lane SW, Williams DA, Watt FM. Modulating the stem cell niche for tissue regeneration. Nat Biotechnol. 2014;32(8):795–803.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Jiang C, Javed A, Kaiser L, Nava MM, Xu R, Brandt DT, et al. Mechanochemical control of epidermal stem cell divisions by B-plexins. Nat Commun. 2021;12(1):1308.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Sedov E, Koren E, Chopra S, Ankawa R, Yosefzon Y, Yusupova M, et al. THY1-mediated mechanisms converge to drive YAP activation in skin homeostasis and repair. Nat Cell Biol. 2022;24(7):1049–63.

    Article  CAS  PubMed  Google Scholar 

  54. Aragona M, Sifrim A, Malfait M, Song Y, Van Herck J, Dekoninck S, et al. Mechanisms of stretch-mediated skin expansion at single-cell resolution. Nature. 2020;584(7820):268–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Xue Y, Lyu C, Taylor A, Van Ee A, Kiemen A, Choi Y, et al. Mechanical tension mobilizes Lgr6(+) epidermal stem cells to drive skin growth. Sci Adv. 2022;8(17):eabl8698.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Carley E, Stewart RM, Zieman A, Jalilian I, King DE, Zubek A, et al. The LINC complex transmits integrin-dependent tension to the nuclear lamina and represses epidermal differentiation. Elife. 2021;1010:e58541.

    Article  Google Scholar 

  57. Ning W, Muroyama A, Li H, Lechler T. Differentiated daughter cells regulate stem cell proliferation and fate through intra-tissue tension. Cell Stem Cell. 2021;28(3):436-452.e5.

    Article  CAS  PubMed  Google Scholar 

  58. Wang M, Zhou X, Zhou S, Wang M, Jiang J, Wu W, et al. Mechanical force drives the initial mesenchymal-epithelial interaction during skin organoid development. Theranostics. 2023;13(9):2930–45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Franco AC, Aveleira C, Cavadas C. Skin senescence: mechanisms and impact on whole-body aging. Trends Mol Med. 2022;28(2):97–109.

    Article  PubMed  Google Scholar 

  60. Doles J, Storer M, Cozzuto L, Roma G, Keyes WM. Age-associated inflammation inhibits epidermal stem cell function. Genes Dev. 2012;26(19):2144–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Dos Santos M, Michopoulou A, André-Frei V, Boulesteix S, Guicher C, Dayan G, et al. Perlecan expression influences the keratin 15-positive cell population fate in the epidermis of aging skin. Aging (Albany NY). 2016;8(4):751–68.

    Article  PubMed  Google Scholar 

  62. Giangreco A, Qin M, Pintar JE, Watt FM. Epidermal stem cells are retained in vivo throughout skin aging. Aging Cell. 2008;7(2):250–9.

    Article  CAS  PubMed  Google Scholar 

  63. Jang H, Jo Y, Lee JH, Choi S. Aging of hair follicle stem cells and their niches. BMB Rep. 2023;56(1):2–9.

    Article  CAS  PubMed  Google Scholar 

  64. Kobielak K, Stokes N, de la Cruz J, Polak L, Fuchs E. Loss of a quiescent niche but not follicle stem cells in the absence of bone morphogenetic protein signaling. Proc Natl Acad Sci U S A. 2007;104(24):10063–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Tiwari RL, Mishra P, Martin N, George NO, Sakk V, Soller K, et al. A Wnt5a-Cdc42 axis controls aging and rejuvenation of hair-follicle stem cells. Aging (Albany NY). 2021;13(4):4778–93.

    Article  CAS  PubMed  Google Scholar 

  66. Li G, Tang X, Zhang S, Jin M, Wang M, Deng Z, et al. SIRT7 activates quiescent hair follicle stem cells to ensure hair growth in mice. Embo j. 2020;39(18):e104365.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Liu Y, Ho C, Wen D, Sun J, Huang L, Gao Y, et al. Targeting the stem cell niche: role of collagen XVII in skin aging and wound repair. Theranostics. 2022;12(15):6446–54.

    Article  PubMed  PubMed Central  Google Scholar 

  68. Liu N, Matsumura H, Kato T, Ichinose S, Takada A, Namiki T, et al. Stem cell competition orchestrates skin homeostasis and ageing. Nature. 2019;568(7752):344–50.

    Article  CAS  PubMed  Google Scholar 

  69. Matsumura H, Mohri Y, Binh NT, Morinaga H, Fukuda M, Ito M, et al. Hair follicle aging is driven by transepidermal elimination of stem cells via COL17A1 proteolysis. Science. 2016;351(6273):aad4395.

    Article  PubMed  Google Scholar 

  70. Zhang C, Wang D, Wang J, Wang L, Qiu W, Kume T, et al. Escape of hair follicle stem cells causes stem cell exhaustion during aging. Nat Aging. 2021;1(10):889–903.

    Article  PubMed  PubMed Central  Google Scholar 

  71. Ge Y, Miao Y, Gur-Cohen S, Gomez N, Yang H, Nikolova M, et al. The aging skin microenvironment dictates stem cell behavior. Proc Natl Acad Sci U S A. 2020;117(10):5339–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Ichijo R, Maki K, Kabata M, Murata T, Nagasaka A, Ishihara S, et al. Vasculature atrophy causes a stiffened microenvironment that augments epidermal stem cell differentiation in aged skin. Nat Aging. 2022;2(7):592–600.

    Article  CAS  PubMed  Google Scholar 

  73. Koester J, Miroshnikova YA, Ghatak S, Chacón-Martínez CA, Morgner J, Li X, et al. Niche stiffening compromises hair follicle stem cell potential during ageing by reducing bivalent promoter accessibility. Nat Cell Biol. 2021;23(7):771–81.

    Article  CAS  PubMed  Google Scholar 

  74. Shin W, Rosin NL, Sparks H, Sinha S, Rahmani W, Sharma N, et al. Dysfunction of hair follicle mesenchymal progenitors contributes to age-associated hair loss. Dev Cell. 2020;53(2):185-198.e7.

    Article  CAS  PubMed  Google Scholar 

  75. Iriyama S, Yasuda M, Nishikawa S, Takai E, Hosoi J, Amano S. Decrease of laminin-511 in the basement membrane due to photoaging reduces epidermal stem/progenitor cells. Sci Rep. 2020;10(1):12592.

    Article  PubMed  PubMed Central  Google Scholar 

  76. Zou Z, Long X, Zhao Q, Zheng Y, Song M, Ma S, et al. A single-cell transcriptomic atlas of human skin aging. Dev Cell. 2021;56(3):383-397.e8.

    Article  CAS  PubMed  Google Scholar 

  77. Chen CC, Murray PJ, Jiang TX, Plikus MV, Chang YT, Lee OK, et al. Regenerative hair waves in aging mice and extra-follicular modulators follistatin, dkk1, and sfrp4. J Invest Dermatol. 2014;134(8):2086–96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Andersen B, Duan J, Karri SS. How and why the circadian clock regulates proliferation of adult epithelial stem cells. Stem Cells. 2023;41(4):319–27.

    Article  PubMed  PubMed Central  Google Scholar 

  79. Niu Y, Wang Y, Chen H, Liu X, Liu J. Overview of the circadian clock in the hair follicle cycle. Biomolecules. 2023. https://doi.org/10.3390/biom13071068.

    Article  PubMed  PubMed Central  Google Scholar 

  80. Benitah SA, Welz PS. Circadian regulation of adult stem cell homeostasis and aging. Cell Stem Cell. 2020;26(6):817–31.

    Article  CAS  PubMed  Google Scholar 

  81. Geyfman M, Kumar V, Liu Q, Ruiz R, Gordon W, Espitia F, et al. Brain and muscle Arnt-like protein-1 (BMAL1) controls circadian cell proliferation and susceptibility to UVB-induced DNA damage in the epidermis. Proc Natl Acad Sci U S A. 2012;109(29):11758–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Wang H, van Spyk E, Liu Q, Geyfman M, Salmans ML, Kumar V, et al. Time-restricted feeding shifts the skin circadian clock and alters UVB-induced DNA damage. Cell Rep. 2017;20(5):1061–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Silveira EJD, Nascimento Filho CHV, Yujra VQ, Webber LP, Castilho RM, Squarize CH. BMAL1 modulates epidermal healing in a process involving the antioxidative defense mechanism. Int J Mol Sci. 2020. https://doi.org/10.3390/ijms21030901.

    Article  PubMed  PubMed Central  Google Scholar 

  84. Janich P, Toufighi K, Solanas G, Luis NM, Minkwitz S, Serrano L, et al. Human epidermal stem cell function is regulated by circadian oscillations. Cell Stem Cell. 2013;13(6):745–53.

    Article  CAS  PubMed  Google Scholar 

  85. Kang TH, Lindsey-Boltz LA, Reardon JT, Sancar A. Circadian control of XPA and excision repair of cisplatin-DNA damage by cryptochrome and HERC2 ubiquitin ligase. Proc Natl Acad Sci U S A. 2010;107(11):4890–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Fan SM, Chang YT, Chen CL, Wang WH, Pan MK, Chen WP, et al. External light activates hair follicle stem cells through eyes via an ipRGC-SCN-sympathetic neural pathway. Proc Natl Acad Sci U S A. 2018;115(29):E6880-e6889.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Welz PS, Zinna VM, Symeonidi A, Koronowski KB, Kinouchi K, Smith JG, et al. BMAL1-driven tissue clocks respond independently to light to maintain homeostasis. Cell. 2019;177(6):1436-1447.e12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Janich P, Pascual G, Merlos-Suárez A, Batlle E, Ripperger J, Albrecht U, et al. The circadian molecular clock creates epidermal stem cell heterogeneity. Nature. 2011;480(7376):209–14.

    Article  CAS  PubMed  Google Scholar 

  89. Kazama I. Stabilizing mast cells by commonly used drugs: a novel therapeutic target to relieve post-COVID syndrome? Drug Discov Ther. 2020;14(5):259–61.

    Article  CAS  PubMed  Google Scholar 

  90. Veltri A, Lang CMR, Cangiotti G, Chan CK, Lien WH. ROR2 regulates self-renewal and maintenance of hair follicle stem cells. Nat Commun. 2022;13(1):4449.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Gonzalez-Fernandez C, González P, Rodríguez FJ. New insights into Wnt signaling alterations in amyotrophic lateral sclerosis: a potential therapeutic target? Neural Regen Res. 2020;15(9):1580–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Rishikaysh P, Dev K, Diaz D, Qureshi WM, Filip S, Mokry J. Signaling involved in hair follicle morphogenesis and development. Int J Mol Sci. 2014;15(1):1647–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Taipale J, Beachy PA. The Hedgehog and Wnt signalling pathways in cancer. Nature. 2001;411(6835):349–54.

    Article  CAS  PubMed  Google Scholar 

  94. Hu XM, Li ZX, Zhang DY, Yang YC, Fu SA, Zhang ZQ, et al. A systematic summary of survival and death signalling during the life of hair follicle stem cells. Stem Cell Res Ther. 2021;12(1):453.

    Article  PubMed  PubMed Central  Google Scholar 

  95. Morinaga H, Mohri Y, Grachtchouk M, Asakawa K, Matsumura H, Oshima M, et al. Obesity accelerates hair thinning by stem cell-centric converging mechanisms. Nature. 2021;595(7866):266–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Wang GY, Wang J, Mancianti ML, Epstein EH Jr. Basal cell carcinomas arise from hair follicle stem cells in Ptch1(+/-) mice. Cancer Cell. 2011;19(1):114–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Artavanis-Tsakonas S, Rand MD, Lake RJ. Notch signaling: cell fate control and signal integration in development. Science. 1999;284(5415):770–6.

    Article  CAS  PubMed  Google Scholar 

  98. Lin MH, Leimeister C, Gessler M, Kopan R. Activation of the Notch pathway in the hair cortex leads to aberrant differentiation of the adjacent hair-shaft layers. Development. 2000;127(11):2421–32.

    Article  CAS  PubMed  Google Scholar 

  99. Lu Z, Xie Y, Huang H, Jiang K, Zhou B, Wang F, et al. Hair follicle stem cells regulate retinoid metabolism to maintain the self-renewal niche for melanocyte stem cells. Elife. 2020;9:e52712.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Shi Y, Massagué J. Mechanisms of TGF-beta signaling from cell membrane to the nucleus. Cell. 2003;113(6):685–700.

    Article  CAS  PubMed  Google Scholar 

  101. Adam RC, Yang H, Ge Y, Lien WH, Wang P, Zhao Y, et al. Temporal layering of signaling effectors drives chromatin remodeling during hair follicle stem cell lineage progression. Cell Stem Cell. 2018;22(3):398-413.e7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Infarinato NR, Stewart KS, Yang Y, Gomez NC, Pasolli HA, Hidalgo L, et al. BMP signaling: at the gate between activated melanocyte stem cells and differentiation. Genes Dev. 2020;34(23–24):1713–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Kobielak K, Pasolli HA, Alonso L, Polak L, Fuchs E. Defining BMP functions in the hair follicle by conditional ablation of BMP receptor IA. J Cell Biol. 2003;163(3):609–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Zhao X, Bian R, Wang F, Wang Y, Li X, Guo Y, et al. GDF-5 promotes epidermal stem cells proliferation via Foxg1-cyclin D1 signaling. Stem Cell Res Ther. 2021;12(1):42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Wagner RN, Piñón Hofbauer J, Wally V, Kofler B, Schmuth M, De Rosa L, et al. Epigenetic and metabolic regulation of epidermal homeostasis. Exp Dermatol. 2021;30(8):1009–22.

    Article  PubMed  PubMed Central  Google Scholar 

  106. Ito K, Ito K. Metabolism and the control of cell fate decisions and stem cell renewal. Annu Rev Cell Dev Biol. 2016;32:399–409.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Kobayashi CI, Suda T. Regulation of reactive oxygen species in stem cells and cancer stem cells. J Cell Physiol. 2012;227(2):421–30.

    Article  CAS  PubMed  Google Scholar 

  108. Kim CS, Ding X, Allmeroth K, Biggs LC, Kolenc OI, L’Hoest N, et al. Glutamine metabolism controls stem cell fate reversibility and long-term maintenance in the hair follicle. Cell Metab. 2020;32(4):629-642 e8.

    Article  CAS  PubMed  Google Scholar 

  109. Wang G, Sweren E, Andrews W, Li Y, Chen J, Xue Y, et al. Commensal microbiome promotes hair follicle regeneration by inducing keratinocyte HIF-1α signaling and glutamine metabolism. Sci Adv. 2023;9(1):eabo7555.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Vietri Rudan M, Mishra A, Klose C, Eggert US, Watt FM. Human epidermal stem cell differentiation is modulated by specific lipid subspecies. Proc Natl Acad Sci U S A. 2020;117(36):22173–82.

    Article  PubMed  PubMed Central  Google Scholar 

  111. White AC, Lowry WE. Refining the role for adult stem cells as cancer cells of origin. Trends Cell Biol. 2015;25(1):11–20.

    Article  CAS  PubMed  Google Scholar 

  112. Baksh SC, Todorova PK, Gur-Cohen S, Hurwitz B, Ge Y, Novak JSS, et al. Extracellular serine controls epidermal stem cell fate and tumour initiation. Nat Cell Biol. 2020;22(7):779–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Tierney MT, Polak L, Yang Y, Abdusselamoglu MD, Baek I, Stewart KS, et al. Vitamin A resolves lineage plasticity to orchestrate stem cell lineage choices. Science. 2024;383(6687):eadi7342.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Lee JH, Choi S. Deciphering the molecular mechanisms of stem cell dynamics in hair follicle regeneration. Exp Mol Med. 2024;56(1):110–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Liu S, Hur YH, Cai X, Cong Q, Yang Y, Xu C, et al. A tissue injury sensing and repair pathway distinct from host pathogen defense. Cell. 2023;186(10):2127-2143 e22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Cheng D, Zhu X, Yan S, Shi L, Liu Z, Zhou X, et al. New insights into inflammatory memory of epidermal stem cells. Front Immunol. 2023;14:1188559.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Gurtner GC, Werner S, Barrandon Y, Longaker MT. Wound repair and regeneration. Nature. 2008;453(7193):314–21.

    Article  CAS  PubMed  Google Scholar 

  118. Xiao T, Yan Z, Xiao S, Xia Y. Proinflammatory cytokines regulate epidermal stem cells in wound epithelialization. Stem Cell Res Ther. 2020;11(1):232.

    Article  PubMed  PubMed Central  Google Scholar 

  119. Yang G, Chen H, Chen Q, Qiu J, Qahar M, Fan Z, et al. Injury-induced interleukin-1 alpha promotes Lgr5 hair follicle stem cells de novo regeneration and proliferation via regulating regenerative microenvironment in mice. Inflamm Regen. 2023;43(1):14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Villarreal-Ponce A, Tiruneh MW, Lee J, Guerrero-Juarez CF, Kuhn J, David JA, et al. Keratinocyte-macrophage crosstalk by the Nrf2/Ccl2/EGF signaling axis orchestrates tissue repair. Cell Rep. 2020;33(8):108417.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Yang R, Liu F, Wang J, Chen X, Xie J, Xiong K. Epidermal stem cells in wound healing and their clinical applications. Stem Cell Res Ther. 2019;10(1):229.

    Article  PubMed  PubMed Central  Google Scholar 

  122. Rousselle P, Braye F, Dayan G. Re-epithelialization of adult skin wounds: cellular mechanisms and therapeutic strategies. Adv Drug Deliv Rev. 2019;146:344–65.

    Article  CAS  PubMed  Google Scholar 

  123. Holt JR, Zeng WZ, Evans EL, Woo SH, Ma S, Abuwarda H, et al. Spatiotemporal dynamics of PIEZO1 localization controls keratinocyte migration during wound healing. Elife. 2021;10:e65415.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Roshan A, Murai K, Fowler J, Simons BD, Nikolaidou-Neokosmidou V, Jones PH. Human keratinocytes have two interconvertible modes of proliferation. Nat Cell Biol. 2016;18(2):145–56.

    Article  CAS  PubMed  Google Scholar 

  125. Sun X, Joost S, Kasper M. Plasticity of Epithelial Cells during Skin Wound Healing. Cold Spring Harb Perspect Biol. 2023;15(5):a041232.

    Article  CAS  PubMed  Google Scholar 

  126. Joost S, Jacob T, Sun X, Annusver K, La Manno G, Sur I, et al. Single-cell transcriptomics of traced epidermal and hair follicle stem cells reveals rapid adaptations during wound healing. Cell Rep. 2018;25(3):585-597.e7.

    Article  CAS  PubMed  Google Scholar 

  127. Adam RC, Yang H, Ge Y, Infarinato NR, Gur-Cohen S, Miao Y, et al. NFI transcription factors provide chromatin access to maintain stem cell identity while preventing unintended lineage fate choices. Nat Cell Biol. 2020;22(6):640–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Ge Y, Gomez NC, Adam RC, Nikolova M, Yang H, Verma A, et al. Stem cell lineage infidelity drives wound repair and cancer. Cell. 2017;169(4):636-650.e14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Gonzales KAU, Polak L, Matos I, Tierney MT, Gola A, Wong E, et al. Stem cells expand potency and alter tissue fitness by accumulating diverse epigenetic memories. Science. 2021;374(6571):eabh2444.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Huang S, Hu Z, Wang P, Zhang Y, Cao X, Dong Y, et al. Rat epidermal stem cells promote the angiogenesis of full-thickness wounds. Stem Cell Res Ther. 2020;11(1):344.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Liu Y, Yang S, Zeng Y, Tang Z, Zong X, Li X, et al. Dysregulated behaviour of hair follicle stem cells triggers alopecia and provides potential therapeutic targets. Exp Dermatol. 2022;31(7):986–92.

    Article  PubMed  Google Scholar 

  132. Zheng W, Xu CH. Innovative approaches and advances for hair follicle regeneration. ACS Biomater Sci Eng. 2023;9(5):2251–76.

    Article  CAS  PubMed  Google Scholar 

  133. Harries M, Hardman J, Chaudhry I, Poblet E, Paus R. Profiling the human hair follicle immune system in lichen planopilaris and frontal fibrosing alopecia: can macrophage polarization differentiate these two conditions microscopically? Br J Dermatol. 2020;183(3):537–47.

    Article  CAS  PubMed  Google Scholar 

  134. Rongioletti F, Christana K. Cicatricial (scarring) alopecias: an overview of pathogenesis, classification, diagnosis, and treatment. Am J Clin Dermatol. 2012;13(4):247–60.

    Article  PubMed  Google Scholar 

  135. Griffiths CEM, Armstrong AW, Gudjonsson JE, Barker J. Psoriasis. Lancet. 2021;397(10281):1301–15.

    Article  CAS  PubMed  Google Scholar 

  136. Eyermann CE, Chen X, Somuncu OS, Li J, Joukov AN, Chen J, et al. ΔNp63 regulates homeostasis, stemness, and suppression of inflammation in the adult epidermis. J Invest Dermatol. 2024;144(1):73-83.e10.

    Article  CAS  PubMed  Google Scholar 

  137. Gago-Lopez N, Mellor LF, Megías D, Martín-Serrano G, Izeta A, Jimenez F, et al. Role of bulge epidermal stem cells and TSLP signaling in psoriasis. EMBO Mol Med. 2019;11(11):e10697.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Augustin I, Gross J, Baumann D, Korn C, Kerr G, Grigoryan T, et al. Loss of epidermal Evi/Wls results in a phenotype resembling psoriasiform dermatitis. J Exp Med. 2013;210(9):1761–77.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Amberg N, Holcmann M, Stulnig G, Sibilia M. Effects of imiquimod on hair follicle stem cells and hair cycle progression. J Invest Dermatol. 2016;136(11):2140–9.

    Article  CAS  PubMed  Google Scholar 

  140. Larsen SB, Cowley CJ, Sajjath SM, Barrows D, Yang Y, Carroll TS, et al. Establishment, maintenance, and recall of inflammatory memory. Cell Stem Cell. 2021;28(10):1758-1774.e8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Naik S, Larsen SB, Gomez NC, Alaverdyan K, Sendoel A, Yuan S, et al. Inflammatory memory sensitizes skin epithelial stem cells to tissue damage. Nature. 2017;550(7677):475–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Niec RE, Rudensky AY, Fuchs E. Inflammatory adaptation in barrier tissues. Cell. 2021;184(13):3361–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Al-Refu K, Goodfield M. Hair follicle stem cells in the pathogenesis of the scarring process in cutaneous lupus erythematosus. Autoimmun Rev. 2009;8(6):474–7.

    Article  PubMed  Google Scholar 

  144. Al-Refu K, Edward S, Ingham E, Goodfield M. Expression of hair follicle stem cells detected by cytokeratin 15 stain: implications for pathogenesis of the scarring process in cutaneous lupus erythematosus. Br J Dermatol. 2009;160(6):1188–96.

    Article  CAS  PubMed  Google Scholar 

  145. Matusiak Ł. Profound consequences of hidradenitis suppurativa: a review. Br J Dermatol. 2020;183(6):e171–7.

    Article  CAS  PubMed  Google Scholar 

  146. Orvain C, Lin YL, Jean-Louis F, Hocini H, Hersant B, Bennasser Y, et al. Hair follicle stem cell replication stress drives IFI16/STING-dependent inflammation in hidradenitis suppurativa. J Clin Invest. 2020;130(7):3777–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Touni AA, Shivde RS, Echuri H, Abdel-Aziz RTA, Abdel-Wahab H, Kundu RV, et al. Melanocyte-keratinocyte cross-talk in vitiligo. Front Med (Lausanne). 2023;10:1176781.

    Article  PubMed  Google Scholar 

  148. Boukhedouni N, Martins C, Darrigade AS, Drullion C, Rambert J, Barrault C, et al. Type-1 cytokines regulate MMP-9 production and E-cadherin disruption to promote melanocyte loss in vitiligo. JCI Insight. 2020;5(11):e133772.

    PubMed  PubMed Central  Google Scholar 

  149. Kovacs D, Bastonini E, Briganti S, Ottaviani M, D’Arino A, Truglio M, et al. Altered epidermal proliferation, differentiation, and lipid composition: novel key elements in the vitiligo puzzle. Sci Adv. 2022;8(35):eabn9299.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Banks-Schlegel S, Green H. Formation of epidermis by serially cultivated human epidermal cells transplanted as an epithelium to athymic mice. Transplantation. 1980;29(4):308–13.

    Article  CAS  PubMed  Google Scholar 

  151. Rheinwald JG, Green H. Formation of a keratinizing epithelium in culture by a cloned cell line derived from a teratoma. Cell. 1975;6(3):317–30.

    Article  CAS  PubMed  Google Scholar 

  152. Rheinwald JG, Green H. Serial cultivation of strains of human epidermal keratinocytes: the formation of keratinizing colonies from single cells. Cell. 1975;6(3):331–43.

    Article  CAS  PubMed  Google Scholar 

  153. Hynds RE, Bonfanti P, Janes SM. Regenerating human epithelia with cultured stem cells: feeder cells, organoids and beyond. EMBO Mol Med. 2018;10(2):139–50.

    Article  CAS  PubMed  Google Scholar 

  154. Wang M, Wu P, Huang J, Liu W, Qian H, Sun Y, et al. Skin cell-derived extracellular vesicles: a promising therapeutic strategy for cutaneous injury. Burns Trauma. 2022;10:tkac037.

    Article  PubMed  PubMed Central  Google Scholar 

  155. Yuan AR, Bian Q, Gao JQ. Current advances in stem cell-based therapies for hair regeneration. Eur J Pharmacol. 2020;881:173197.

    Article  CAS  PubMed  Google Scholar 

  156. Zhang Q, Wen J, Liu C, Ma C, Bai F, Leng X, et al. Early-stage bilayer tissue-engineered skin substitute formed by adult skin progenitor cells produces an improved skin structure in vivo. Stem Cell Res Ther. 2020;11(1):407.

    Article  PubMed  PubMed Central  Google Scholar 

  157. Jahoda CA, Horne KA, Oliver RF. Induction of hair growth by implantation of cultured dermal papilla cells. Nature. 1984;311(5986):560–2.

    Article  CAS  PubMed  Google Scholar 

  158. Wang X, Wang X, Liu J, Cai T, Guo L, Wang S, et al. Hair follicle and sebaceous gland de novo regeneration with cultured epidermal stem cells and skin-derived precursors. Stem Cells Transl Med. 2016;5(12):1695–706.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Lee J, Rabbani CC, Gao H, Steinhart MR, Woodruff BM, Pflum ZE, et al. Hair-bearing human skin generated entirely from pluripotent stem cells. Nature. 2020;582(7812):399–404.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Shafiee A, Sun J, Ahmed IA, Phua F, Rossi GR, Lin CY, et al. Development of physiologically relevant skin organoids from human induced pluripotent stem cells. Small. 2024;20(16):e2304879.

    Article  PubMed  Google Scholar 

  161. Jayarajan V, Hall GT, Xenakis T, Bulstrode N, Moulding D, Castellano S, et al. Short-term treatment with rho-associated kinase inhibitor preserves keratinocyte stem cell characteristics in vitro. Cells. 2023. https://doi.org/10.3390/cells12030346.

    Article  PubMed  PubMed Central  Google Scholar 

  162. Chen P, Miao Y, Zhang F, Huang J, Chen Y, Fan Z, et al. Nanoscale microenvironment engineering based on layer-by-layer self-assembly to regulate hair follicle stem cell fate for regenerative medicine. Theranostics. 2020;10(25):11673–89.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Hong ZX, Zhu ST, Li H, Luo JZ, Yang Y, An Y, et al. Bioengineered skin organoids: from development to applications. Mil Med Res. 2023;10(1):40.

    PubMed  PubMed Central  Google Scholar 

  164. Ong CS, Yesantharao P, Huang CY, Mattson G, Boktor J, Fukunishi T, et al. 3D bioprinting using stem cells. Pediatr Res. 2018;83(1–2):223–31.

    Article  CAS  PubMed  Google Scholar 

  165. Yang R, Yang S, Zhao J, Hu X, Chen X, Wang J, et al. Progress in studies of epidermal stem cells and their application in skin tissue engineering. Stem Cell Res Ther. 2020;11(1):303.

    Article  PubMed  PubMed Central  Google Scholar 

  166. Zhao M, Wang J, Zhang J, Huang J, Luo L, Yang Y, et al. Functionalizing multi-component bioink with platelet-rich plasma for customized in-situ bilayer bioprinting for wound healing. Mater Today Bio. 2022;16:100334.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  167. Zhao P, Dang Z, Liu M, Guo D, Luo R, Zhang M, et al. Molecular hydrogen promotes wound healing by inducing early epidermal stem cell proliferation and extracellular matrix deposition. Inflamm Regen. 2023;43(1):22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Hirose T, Kotoku J, Toki F, Nishimura EK, Nanba D. Label-free quality control and identification of human keratinocyte stem cells by deep learning-based automated cell tracking. Stem Cells. 2021;39(8):1091–100.

    Article  CAS  PubMed  Google Scholar 

  169. Zhao X, Li X, Wang Y, Guo Y, Huang Y, Lv D, et al. Stability and biosafety of human epidermal stem cell for wound repair: preclinical evaluation. Stem Cell Res Ther. 2023;14(1):4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Wang Z, Xu H, Yang H, Zhang Y, Wang X, Wang P, et al. Single-stage transplantation combined with epidermal stem cells promotes the survival of tissue-engineered skin by inducing early angiogenesis. Stem Cell Res Ther. 2023;14(1):51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Kwak S, Song CL, Lee J, Kim S, Nam S, Park YJ, et al. Development of pluripotent stem cell-derived epidermal organoids that generate effective extracellular vesicles in skin regeneration. Biomaterials. 2024;307:122522.

    Article  CAS  PubMed  Google Scholar 

  172. Leng L, Ma J, Lv L, Wang W, Gao D, Zhu Y, et al. Both Wnt signaling and epidermal stem cell-derived extracellular vesicles are involved in epidermal cell growth. Stem Cell Res Ther. 2020;11(1):415.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Niti A, Koliakos G, Michopoulou A. Stem cell therapies for epidermolysis bullosa treatment. Bioengineering (Basel). 2023;10(4):422.

    Article  CAS  PubMed  Google Scholar 

  174. De Rosa L, Enzo E, Palamenghi M, Sercia L, De Luca M. Stairways to advanced therapies for epidermolysis bullosa. Cold Spring Harb Perspect Biol. 2023;15(4):a041229.

    Article  PubMed  Google Scholar 

  175. De Rosa L, Enzo E, Zardi G, Bodemer C, Magnoni C, Schneider H, et al. Hologene 5: a phase II/III clinical trial of combined cell and gene therapy of junctional epidermolysis bullosa. Front Genet. 2021;12:705019.

    Article  PubMed  PubMed Central  Google Scholar 

  176. Mavilio F, Pellegrini G, Ferrari S, Di Nunzio F, Di Iorio E, Recchia A, et al. Correction of junctional epidermolysis bullosa by transplantation of genetically modified epidermal stem cells. Nat Med. 2006;12(12):1397–402.

    Article  CAS  PubMed  Google Scholar 

  177. Hirsch T, Rothoeft T, Teig N, Bauer JW, Pellegrini G, De Rosa L, et al. Regeneration of the entire human epidermis using transgenic stem cells. Nature. 2017;551(7680):327–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Kueckelhaus M, Rothoeft T, De Rosa L, Yeni B, Ohmann T, Maier C, et al. Transgenic epidermal cultures for junctional epidermolysis bullosa—5-year outcomes. N Engl J Med. 2021;385(24):2264–70.

    Article  PubMed  Google Scholar 

  179. De Rosa L, Secone Seconetti A, De Santis G, Pellacani G, Hirsch T, Rothoeft T, et al. Laminin 332-dependent YAP dysregulation depletes epidermal stem cells in junctional epidermolysis bullosa. Cell Rep. 2019;27(7):2036-2049.e6.

    Article  PubMed  Google Scholar 

  180. Eichstadt S, Barriga M, Ponakala A, Teng C, Nguyen NT, Siprashvili Z, et al. Phase 1/2a clinical trial of gene-corrected autologous cell therapy for recessive dystrophic epidermolysis bullosa. JCI Insight. 2019;4(19).

  181. Siprashvili Z, Nguyen NT, Gorell ES, Loutit K, Khuu P, Furukawa LK, et al. Safety and wound outcomes following genetically corrected autologous epidermal grafts in patients with recessive dystrophic epidermolysis bullosa. JAMA. 2016;316(17):1808–17.

    Article  PubMed  Google Scholar 

  182. Kocher T, Bischof J, Haas SA, March OP, Liemberger B, Hainzl S, et al. A non-viral and selection-free COL7A1 HDR approach with improved safety profile for dystrophic epidermolysis bullosa. Mol Ther Nucleic Acids. 2021;25:237–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  183. O’Keeffe Ahern J, Lara-Sáez I, Zhou D, Murillas R, Bonafont J, Mencía Á, et al. Non-viral delivery of CRISPR-Cas9 complexes for targeted gene editing via a polymer delivery system. Gene Ther. 2022;29(3–4):157–70.

    Article  CAS  PubMed  Google Scholar 

  184. Cattaneo C, Enzo E, De Rosa L, Sercia L, Consiglio F, Forcato M, et al. Allele-specific CRISPR-Cas9 editing of dominant epidermolysis bullosa simplex in human epidermal stem cells. Mol Ther. 2024;32(2):372–83.

    Article  CAS  PubMed  Google Scholar 

  185. Kawakami T. Surgical procedures and innovative approaches for vitiligo regenerative treatment and melanocytorrhagy. J Dermatol. 2022;49(4):391–401.

    Article  CAS  PubMed  Google Scholar 

  186. Falabella R, Escobar C, Borrero I. Treatment of refractory and stable vitiligo by transplantation of in vitro cultured epidermal autografts bearing melanocytes. J Am Acad Dermatol. 1992;26(2 Pt 1):230–6.

    Article  CAS  PubMed  Google Scholar 

  187. El-Zawahry BM, Zaki NS, Bassiouny DA, Sobhi RM, Zaghloul A, Khorshied MM, et al. Autologous melanocyte-keratinocyte suspension in the treatment of vitiligo. J Eur Acad Dermatol Venereol. 2011;25(2):215–20.

    Article  CAS  PubMed  Google Scholar 

  188. Komen L, Vrijman C, Tjin EP, Krebbers G, de Rie MA, Luiten RM, et al. Autologous cell suspension transplantation using a cell extraction device in segmental vitiligo and piebaldism patients: a randomized controlled pilot study. J Am Acad Dermatol. 2015;73(1):170–2.

    Article  PubMed  Google Scholar 

  189. Razmi TM, Kumar R, Rani S, Kumaran SM, Tanwar S, Parsad D. Combination of follicular and epidermal cell suspension as a novel surgical approach in difficult-to-treat vitiligo: a randomized clinical trial. JAMA Dermatol. 2018;154(3):301–8.

    Article  Google Scholar 

  190. Jensen KB, Driskell RR, Watt FM. Assaying proliferation and differentiation capacity of stem cells using disaggregated adult mouse epidermis. Nat Protoc. 2010;5(5):898–911.

    Article  CAS  PubMed  Google Scholar 

  191. Li J, Xu X, Tiwari M, Chen Y, Fuller M, Bansal V, et al. SPT6 promotes epidermal differentiation and blockade of an intestinal-like phenotype through control of transcriptional elongation. Nat Commun. 2021;12(1):784.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  192. Yang R, Zheng Y, Burrows M, Liu S, Wei Z, Nace A, et al. Generation of folliculogenic human epithelial stem cells from induced pluripotent stem cells. Nat Commun. 2014;5:3071.

    Article  PubMed  Google Scholar 

  193. Hu Z, Chen Y, Gao M, Chi X, He Y, Zhang C, et al. Novel strategy for primary epithelial cell isolation: Combination of hyaluronidase and collagenase I. Cell Prolif. 2023;56(1):e13320.

    Article  CAS  PubMed  Google Scholar 

  194. Jones PH, Watt FM. Separation of human epidermal stem cells from transit amplifying cells on the basis of differences in integrin function and expression. Cell. 1993;73(4):713–24.

    Article  CAS  PubMed  Google Scholar 

  195. Kumar S, Poojan S, Verma V, Verma MK, Lohani M. Rapid isolation of integrin rich multipotent stem cell pool and reconstruction of mouse epidermis equivalent. Am J Stem Cells. 2014;3(1):27–36.

    CAS  PubMed  PubMed Central  Google Scholar 

  196. Ghio SC, Barbier MA, Doucet EJ, Debbah I, Safoine M, Le-Bel G, et al. A newly developed chemically defined serum-free medium suitable for human primary keratinocyte culture and tissue engineering applications. Int J Mol Sci. 2023;24(3):1821.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  197. Hirano S, Kageyama T, Yamanouchi M, Yan L, Suzuki K, Ebisawa K, et al. Expansion culture of hair follicle stem cells through uniform aggregation in microwell array devices. ACS Biomater Sci Eng. 2023;9(3):1510–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  198. Braun KM, Niemann C, Jensen UB, Sundberg JP, Silva-Vargas V, Watt FM. Manipulation of stem cell proliferation and lineage commitment: visualisation of label-retaining cells in wholemounts of mouse epidermis. Development. 2003;130(21):5241–55.

    Article  CAS  PubMed  Google Scholar 

  199. Cotsarelis G, Sun TT, Lavker RM. Label-retaining cells reside in the bulge area of pilosebaceous unit: implications for follicular stem cells, hair cycle, and skin carcinogenesis. Cell. 1990;61(7):1329–37.

    Article  CAS  PubMed  Google Scholar 

  200. May D, Yun S, Gonzalez DG, Park S, Chen Y, Lathrop E, et al. Live imaging reveals chromatin compaction transitions and dynamic transcriptional bursting during stem cell differentiation in vivo. Elife. 2023;12:e83444.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  201. Tumbar T, Guasch G, Greco V, Blanpain C, Lowry WE, Rendl M, et al. Defining the epithelial stem cell niche in skin. Science. 2004;303(5656):359–63.

    Article  CAS  PubMed  Google Scholar 

  202. Huang S, Kuri P, Aubert Y, Brewster M, Li N, Farrelly O, et al. Lgr6 marks epidermal stem cells with a nerve-dependent role in wound re-epithelialization. Cell Stem Cell. 2021;28(9):1582-1596.e6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  203. Kretzschmar K, Watt FM. Lineage tracing. Cell. 2012;148(1–2):33–45.

    Article  CAS  PubMed  Google Scholar 

  204. Seminowicz DA, Wideman TH, Naso L, Hatami-Khoroushahi Z, Fallatah S, Ware MA, et al. Effective treatment of chronic low back pain in humans reverses abnormal brain anatomy and function. J Neurosci. 2011;31(20):7540–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  205. Hinnant T, Lechler T. Hair follicle stem cells feel the pressure. Cell Stem Cell. 2022;29(1):1–2.

    Article  CAS  PubMed  Google Scholar 

  206. Gaddameedhi S, Selby CP, Kaufmann WK, Smart RC, Sancar A. Control of skin cancer by the circadian rhythm. Proc Natl Acad Sci U S A. 2011;108(46):18790–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  207. Yang Y, Gomez N, Infarinato N, Adam RC, Sribour M, Baek I, et al. The pioneer factor SOX9 competes for epigenetic factors to switch stem cell fates. Nat Cell Biol. 2023;25(8):1185–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  208. Li J, Jiang TX, Hughes MW, Wu P, Yu J, Widelitz RB, et al. Progressive alopecia reveals decreasing stem cell activation probability during aging of mice with epidermal deletion of DNA methyltransferase 1. J Invest Dermatol. 2012;132(12):2681–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  209. Cohen I, Bar C, Liu H, Valdes VJ, Zhao D, Galbo PM Jr, et al. Polycomb complexes redundantly maintain epidermal stem cell identity during development. Genes Dev. 2021;35(5–6):354–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  210. Xia X, Cao G, Sun G, Zhu L, Tian Y, Song Y, et al. GLS1-mediated glutaminolysis unbridled by MALT1 protease promotes psoriasis pathogenesis. J Clin Invest. 2020;130(10):5180–96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  211. Suen WJ, Li ST, Yang LT. Hes1 regulates anagen initiation and hair follicle regeneration through modulation of hedgehog signaling. Stem Cells. 2020;38(2):301–14.

    Article  CAS  PubMed  Google Scholar 

  212. Jaks V, Barker N, Kasper M, van Es JH, Snippert HJ, Clevers H, et al. Lgr5 marks cycling, yet long-lived, hair follicle stem cells. Nat Genet. 2008;40(11):1291–9.

    Article  CAS  PubMed  Google Scholar 

  213. Chen JK, Wiedemann J, Nguyen L, Lin Z, Tahir M, Hui CC, et al. IRX5 promotes DNA damage repair and activation of hair follicle stem cells. Stem Cell Reports. 2023;18(5):1227–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  214. Ren X, Xia W, Xu P, Shen H, Dai X, Liu M, et al. Lgr4 deletion delays the hair cycle and inhibits the activation of hair follicle stem cells. J Invest Dermatol. 2020;140(9):1706-171 e24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  215. Li J, Chen Y, Tiwari M, Bansal V, Sen GL. Regulation of integrin and extracellular matrix genes by HNRNPL is necessary for epidermal renewal. PLoS Biol. 2021;19(9):e3001378.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  216. Li J, Chen Y, Xu X, Jones J, Tiwari M, Ling J, et al. HNRNPK maintains epidermal progenitor function through transcription of proliferation genes and degrading differentiation promoting mRNAs. Nat Commun. 2019;10(1):4198.

    Article  PubMed  PubMed Central  Google Scholar 

  217. Polito MP, Marini G, Fabrizi A, Sercia L, Enzo E, De Luca M. Biochemical role of FOXM1-dependent histone linker H1B in human epidermal stem cells. Cell Death Dis. 2024;15(7):508.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  218. Liu H, Su P, Li Y, Hoover A, Hu S, King SA, et al. VAMP2 controls murine epidermal differentiation and carcinogenesis by regulation of nucleophagy. Dev Cell. 2024.

  219. Pickup ME, Hu A, Patel HJ, Ahmed MI. MicroRNA-148a controls epidermal and hair follicle stem/progenitor cells by modulating the activities of ROCK1 and ELF5. J Invest Dermatol. 2023;143(3):480–4915.

    Article  CAS  PubMed  Google Scholar 

  220. Lin Z, Jin S, Chen J, Li Z, Lin Z, Tang L, et al. Murine interfollicular epidermal differentiation is gradualistic with GRHL3 controlling progression from stem to transition cell states. Nat Commun. 2020;11(1):5434.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  221. Hiratsuka T, Bordeu I, Pruessner G, Watt FM. Regulation of ERK basal and pulsatile activity control proliferation and exit from the stem cell compartment in mammalian epidermis. Proc Natl Acad Sci U S A. 2020;117(30):17796–807.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  222. Liu F, Zhang X, Peng Y, Zhang L, Yu Y, Hua P, et al. miR-24 controls the regenerative competence of hair follicle progenitors by targeting Plk3. Cell Rep. 2021;35(10):109225.

    Article  CAS  PubMed  Google Scholar 

  223. Balmer P, Hariton WVJ, Sayar BS, Jagannathan V, Galichet A, Leeb T, et al. SUV39H2 epigenetic silencing controls fate conversion of epidermal stem and progenitor cells. J Cell Biol. 2021. https://doi.org/10.1083/jcb.201908178.

    Article  PubMed  PubMed Central  Google Scholar 

  224. Wang L, Siegenthaler JA, Dowell RD, Yi R. Foxc1 reinforces quiescence in self-renewing hair follicle stem cells. Science. 2016;351(6273):613–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  225. Lien WH, Polak L, Lin M, Lay K, Zheng D, Fuchs E. In vivo transcriptional governance of hair follicle stem cells by canonical Wnt regulators. Nat Cell Biol. 2014;16(2):179–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This work was supported by the National Natural Science Foundation of China (82273520 and 82322057).

Author information

Authors and Affiliations

Authors

Contributions

X.T., J.W., J.C., G.W., and S.S.: the conceptualization and design of this study. X.T., J.W., and J.C.: the investigation and methodology of this study. X.T., J.W., J.C., W.L., P.Q., H.Q., Z.L., and E.D.: writing original draft. X.T., J.W., and J.C.: Figures and tables of the manuscript. X.T., J.W., J.C., and S.S.: writing-review & editing of the manuscript. S.S.: the funding acquisition. All authors have read and approved the final manuscript.

Corresponding authors

Correspondence to Gang Wang or Shuai Shao.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

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

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tang, X., Wang, J., Chen, J. et al. Epidermal stem cells: skin surveillance and clinical perspective. J Transl Med 22, 779 (2024). https://doi.org/10.1186/s12967-024-05600-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12967-024-05600-1

Keywords