key: cord-104031-vodwptdo authors: Altshuler, Anna; Amitai-Lange, Aya; Tarazi, Noam; Dey, Sunanda; Strinkovsky, Lior; Bhattacharya, Swarnabh; Hadad-Porat, Shira; Nasser, Waseem; Imeri, Jusuf; Ben-David, Gil; Tiosano, Beatrice; Berkowitz, Eran; Karin, Nathan; Savir, Yonatan; Shalom-Feuerstein, Ruby title: Capturing limbal epithelial stem cell population dynamics, signature, and their niche date: 2020-07-01 journal: bioRxiv DOI: 10.1101/2020.06.30.179754 sha: doc_id: 104031 cord_uid: vodwptdo Stem cells (SCs) are traditionally viewed as rare, slow-cycling cells that follow deterministic rules dictating their self-renewal or differentiation. It was several decades ago, when limbal epithelial SCs (LSCs) that regenerate the corneal epithelium were one of the first sporadic, quiescent SCs ever discovered. However, LSC dynamics, heterogeneity and genetic signature are largely unknown. Moreover, recent accumulating evidence strongly suggested that epithelial SCs are actually abundant, frequently dividing cells that display stochastic behavior. In this work, we performed an in-depth analysis of the murine limbal epithelium by single-cell RNA sequencing and quantitative lineage tracing. The generated data provided an atlas of cell states of the corneal epithelial lineage, and particularly, revealed the co-existence of two novel LSC populations that reside in separate and well-defined sub-compartments. In the “outer” limbus, we identified a primitive widespread population of quiescent LSCs (qLSCs) that uniformly express Krt15/Gpha2/Ifitm3/Cd63 proteins, while the “inner” limbus host prevalent active LSCs (aLSCs) co-expressing Krt15-GFP/Atf3/Mt1-2/Socs3. Analysis of LSC population dynamics suggests that while qLSCs and aLSCs possess different proliferation rates, they both follow similar stochastic rules that dictate their self-renewal and differentiation. Finally, T cells were distributed in close proximity to qLSCs. Indeed, their absence or inhibition resulted in the loss of quiescence and delayed wound healing. Taken together, we propose that divergent regenerative strategies are tailored to properly support tissue-specific physiological constraints. The present study suggests that in the case of the cornea, quiescent epithelial SCs are abundant, follow stochastic rules and neutral drift dynamics. Stem cells (SCs) differ from any other cells by their substantial ability to retain in an undifferentiated state, self-duplicate or enter differentiation program, depending on tissue demand 1 2 . SCs have been successfully applied in clinical trials to reconstitute the bone marrow, treat skin burn, and restore corneal blindness 3 . However, various fundamental features of SCs are still under debate or remain unknown. For example, SC prevalence in the niche, SC heterogeneity and self-renewal mechanisms are not well understood, and consequently, they remain enigmatic cells 4 5 . These key challenges hamper the progression towards advanced application of SCs for regenerative medicine. Seminal studies supported a deterministic and hierarchical model in which SCs were viewed as slowcycling cells that are located in a specialized microenvironment known as the niche 6 7 8 . Such quiescent SCs (qSCs) are commonly believed to be rare cells that are surrounded by their progeny, of abundant, fast cycling but short-lived progenitor cells 9 . Key evidence for the quiescence and scarcity of SCs came from indirect observations in vivo and in vitro. Sporadic slow-cycling cells were identified in different tissues as nucleotide-label retaining cells in pulse-chase experiments performed in various tissues including the cornea 10 , bone marrow 11 , brain 12 , skin 13 , gut 14 and skeletal muscle 15 . Clonal survival/growth dynamics in vivo (i.e. by genetic lineage tracing of single cells at the SC compartment) or ex vivo (i.e. by colony formation assay), typically result in the detection of a low number of clones that survive for long-periods of time. This evidence shaped the traditional deterministic model that considered qSCs as very potent cells that are also very scarce. The benefits of quiescence may be paramount and include reduced biochemical damage associated with the production of toxic agents or biosynthesis of macromolecules, and minimalized accumulation of disease-causing mutations 16 17 . Despite much effort in the field, markers that reliably identify genuine and scarce qSCs were usually not found. Proposed markers, typically labeled a contiguous widespread cell population in the SC compartment 5 . Nevertheless, the co-existence of qSCs and actively dividing SCs (aSCs) in specific tissues including the bone marrow, brain, hair follicle and muscle has been shown 18 19 20 . These tissues differ from traditional epithelial SC models by their significantly slower turnover (or resting phase) 21 . In recent years, however, frequently dividing SCs were identified in epithelial tissues including the gut 22 , epidermis 23 and esophagus 24 . These studies proposed an alternative, virtually opposing, stochastic model that attributes SCs with entirely contrary properties including short cell cycle, abundance in the niche and unpredictable survival rate 18 1 25 26 . One of the common models that is used to capture this phenomena is that of asymmetric self-renewal with neural drift dynamics 23 27 . These findings challenged the conventional paradigm that considered quiescence as an integral feature of SCs. It is likely that the appropriate regenerative strategy is tailored for each tissue depending on the physiological necessities and challenges (e.g. irradiation, toxicity, cellular turn over or lineage diversity). In this sense, the corneal epithelium is an interesting study case, whereas SCs and their progeny must support tissue transparency and protection from sun irradiation, toxicants, physical injury and invading microorganisms 28 . Indeed, most ocular surface epithelial cancers are restricted to the corneal SC compartment of the limbus. However, the rareness of limbal tumors suggests the existence of specialized mechanisms that may provide protection against neoplastic cell transformation. The cornea is an excellent model for SC research hallmarked by segregation of SCs, short-lived progenitors and differentiated cells into distinct anatomical compartments. The clarity and high accessibility of the cornea, allows multi-color ("Confetti") fluorescent lineage tracing 29 30 31 , assessing different SC hypotheses 32 , as well as SC or progenitor cell depletion using vital microscopy 33 34 . Corneal epithelial SCs reside in the limbus, a ring-shaped region at the corneal-conjunctival boundary. The limbal microenvironment (niche) markedly differs from that of the cornea. It contains blood and lymph vasculature whereas the cornea is avascular, it is hallmarked by a unique extracellular matrix with soft biochemical properties while it hosts specialized niche cells 35 36 37 . Rare label retaining cells were confined to the limbus 10 38 . In line, colony formation tests revealed that only a low percentage of limbal cells could form large clones ("holoclones") in vitro, a feature that is commonly attributed to bona fide limbal SCs (LSCs) 39 40 . Based on these observations, it is widely believed that true LSCs are scarce limbal epithelial cells kept most of the time in quiescence state and are surrounded by their early progeny which are abundant, fast dividing but short-lived progenitor cells. Tremendous efforts have been made by many research groups to discover markers for the identification of LSCs. Keratin 15 (Krt15) 33 41 , C/EBPdelta and BMI1 42 , ABCB5 43 , ABCG2 44 and P63 45 represent a partial list of genes proposed to identify LSCs. Recently, we have shown that the Krt15-GFP transgene (green fluorescent protein coding gene under the promoter of Krt15) labeled a discrete population of murine LSCs. Krt15-GFP + basal limbal epithelial cells were located at close proximity to the site of corneal regeneration origin, as evident by lineage tracing of Krt14 + cells 33 . However, neither Krt15-GFP, nor any of the proposed markers could faithfully mark a hypothetical rare, slow-cycling SC population in the limbus, suggesting that the current model may be inaccurate. Consequently, the prevalence of LSCs, their genetic signature, and regulation by niche cells remain poorly defined. Here we combined single-cell transcriptomics and quantitative lineage tracing to capture LSC populations and their signature. This work provides a useful atlas of the entire murine corneal epithelial lineage, including the signature of a slow cycling quiescent LSC (qLSC) state, and finally, unravels an apparently tight interaction and regulation of LSCs by T cells. Careful inspection of the limbus of Krt15-GFP transgene bearing mice showed that Krt15-GFP + cells are confined to an "inner" limbal zone, suggesting that the "outer" limbal region (the space between Krt15-GFP + cells and the Krt8 + conjunctiva) contains a different LSC population (Fig. 1a, S1a) . To clarify the molecular nature and heterogeneity of LSC populations, we performed a single-cell transcriptomic analysis. In order to facilitate the detection of rare cell populations, reduce variability and gain robust statistical data, we preferentially isolated epithelial cells from the limbus of 5 individual adult mice (2.5 months old, n=10 eyes). The limbus together with marginal conjunctiva and corneal periphery were carefully dissected (Fig. 1a) , a protocol for isolating mainly epithelial cells was applied (see Methods), and suspended cells were subjected to 10x Chromium Single-Cell RNA sequencing (scRNA-seq). Initial raw data analysis and quality controls were performed with Cell Ranger software and a large number of 4,787 high quality cells passed rigorous quality tests exhibiting significant mean reads (98,763 per cell) as well as a substantial median number of 2,479 detected genes per cell. In silico analysis revealed discrete cell states in the corneal epithelial lineage Cell filtering and unbiased clustering were performed with R package Seurat 46 revealing 11 cell populations displaying a markedly distinct signature of gene expression (Fig. 1b, S1b) . Analysis of lineagespecific markers suggested that >99% of the cells are ocular epithelial cells (see Methods). Next, we performed in silico analysis of putative markers for each epithelial tissue (i.e. conjunctiva, limbus, and cornea) and cell layer (i.e. basal or supra basal layer) ( Fig. 1c -j, S1c-d). Conjunctival cells: Clusters 1-2 were identified as conjunctival cells based on the positive expression of putative conjunctival cytokeratins (Krt17, Krt4, Krt19 (Fig. 1c) and Krt6a, Krt13, Krt8 (Fig. S1c) 47 ), and the lack or comparatively low expression of corneal markers Krt12 and Slurp1 48 49 and of the protein phosphatase 1 (Ppp1r3c) (Fig. 1d ). Further analysis of putative markers of basal (Itgb1, Itgb4, Ccnd1 (Cyclin D1) 50 51 ) and supra basal (Cld4, Dsg1a, Cdkn1a (P21) 52 53 54 ) cells indicated that cluster 1 represents a population of basal conjunctival cells while cluster 2 denotes supra-basal conjunctival cells (Fig. 1g-h) . Krt15 was highly expressed by conjunctival basal cells (cluster 1) and even higher in conjunctival suprabasal cells (cluster 2), whereas Krt14 displayed basal cell enrichment in conjunctival cells (Fig. 1i) . Limbal cells: Clusters 3-5 were identified as limbal cells based on the lack or lower expression of conjunctival (Fig. 1c , S1c) and corneal (Fig. 1d ) markers. While cluster 5 displayed supra-basal cell phenotype, cluster 3-4 were identified as limbal basal cells (Fig. 1g -h) expressing a set of new limbal specific markers (Fig. 1e-f ). As compared to cluster 4, cluster 3 was hallmarked by much higher levels of Krt15 and Krt14 (Fig. 1i ) and lower Krt12 (Fig. 1d) , suggesting that it represents the most primitive undifferentiated limbal cell population. Corneal cells: Although a minimal basal expression levels of corneal specific Krt12 mRNA was found in all groups, clusters 6-9 expressed much higher levels of Krt12 (Krt12 hi ) (Fig. 1d) . Clusters 6-7 were basal corneal epithelial markers although cluster 7 appeared to include cells that seemingly initiated differentiation, as evident by attenuation of basal epithelial cell markers on the expense of supra-basal cell markers ( Fig. 1g-h) . Likewise, analysis of basal/supra-basal cell markers suggested that cluster 8 represents partially differentiated cells, most likely corneal wing cells, whereas cluster 9 represents terminally differentiated superficial cells (Fig. 1g-h) . In line with previous reports 33 41 , while Krt15 and Krt15-GFP labeled basal conjunctival/limbal epithelial cells, they also marked supra-basal cells of the conjunctiva, limbus and marginal cornea periphery (Fig. 1i) . Interestingly, the signal of endogenous Krt15 did not overlap with that of Krt15-GFP (Fig. 1l ), in line with previous reports on epidermis 55 . This suggests that the genomic insertion site of K15-GFP transgene, the fractional Krt15 promoter cloned, lack of Krt15 coding sequence or presence of GFP sequence, may have deviated the transcriptional regulation of the Krt15-GFP transgene away from that of endogenous Krt15. Cells in mitosis: Interestingly, clusters 10-11 were found to represent cells from all 3 tissues (conjunctiva, limbus, cornea (Figs. 1c-f)) that were hallmarked by various cell cycle genes including Mki67, Top2a, Ccna2 (Cyclin A2) (Fig. 1j) . The grouping of this cell population of mixed lineages into discrete clusters indicated that the changes in expression of genes related to mitosis were profound and dominated the differential expression of tissue-specific genes. Further analysis of cell cycle genes, suggested that cluster 10 represents cells that were captured at early stage of mitosis (late G1-S) while cells of cluster 11 were netted at relatively more advanced stages (S, G2 and M) of the cell cycle (Fig. S1d) . Two novel limbal sub-compartments: To certify the cluster identification in vivo, we performed in situ hybridization (ISH) on tissue sections to assess the anatomical localization of selected clusters. Eyes of 2-3 months old C57BL/6 mice were enucleated and tissue sections were stained as detailed in Methods. The conjunctiva and cornea were labeled using ISH probes for Keratin 4 (Krt4) and Krt12, respectively. In order to identify the location of the limbal clusters 3-4, we focused on Gpha2 and Atf3 that displayed relatively high mRNA copies. Interestingly, this analysis revealed 2 novel well-demarcated sub-compartments in the limbus. The "outer limbus" was occupied by thin epithelium comprised of small basal epithelial cells with flattened morphology that expressed Gpha2 mRNA. The basal layer cells of the "inner limbus" expressed Atf3 mRNA and displayed flattened cell morphology that gradually became cuboidal towards the corneal periphery (Fig. 1k) . In agreement, immunostaining showed that the outer limbal basal epithelial cells were Krt15 + /Ifitm3 + /Cd63 + whereas inner limbal epithelial cells were Krt15-GFP + /Atf3 + /Mt1-2 + (Fig. 1l) . It was noticeable that the two limbal sub-compartments are thin. Moreover, the outer limbus tissue was occasionally absent or truncated in tissue sections. Therefore, to keep the tissue intact and facilitate efficient analysis, we established wholemount immunostaining for antibodies that were found suitable. Figure 1m -o, shows that antibodies against Gpha2, Ifitm3 and Cd63 specifically labeled the outer limbus while Krt15-GFP and Mt1 labeled the inner limbus by wholemount immunostaining. For consistency and further analyses, we defined conventions for the identification of the outer and inner limbus. The outer limbus markers (Gpha2, Ifitm3 or Cd63) displayed uniform and sharp signal and were well-correlated with slow proliferation and clonal growth dynamics (Figs 2-3) . Hence, these markers hereafter were chosen to define the outer limbus, as illustrated in Fig. 1o . Next, the inner limbus was defined as the limbal zone that was negative to outer limbal markers, and its boundary with the peripheral cornea was demarcated by inner limbus markers (e.g. Krt15-GFP in Fig. 1o ). Accordingly, quantitative analysis showed that the average width of the outer limbus was ~100m while the inner limbus was much wider (~240m) (Fig. 1p) . Interestingly, the pattern of Krt15-GFP signal was often continuous, or nearly continuous with outer limbus markers like Gpha2, but never overlapped with the labeling of the outer limbus ( Fig. 1n-o) . Interestingly, by microscopic analysis of wholemount tissues, it was noticed that flattened basal cells found in both sub-compartments were hallmarked by a unique pattern of nuclei staining that differed from that of the cuboidal basal cells that were detected in the marginal inner limbus and corneal periphery ( Fig. S1e-f ). Finally, analysis of selected keratin coding genes confirmed that basal outer limbal epithelial cells were Krt15 hi /Krt14 hi /Krt17 med /Krt12 neg , basal inner limbal epithelial cells were Krt15 neg /Krt14 med /Krt17 neg /Krt12 low , while conjunctival cells were Krt19 hi /Krt17 hi /Krt6 hi /Krt8 med /Krt12 neg (Fig. S1c, g-h) . In silico analysis of clusters: Next, pathway enrichment analysis using WebGestalt 56 was performed with the ORA method using KEGG as the functional database in order to identify potential pathways that were enriched in clusters 3 or 4 in comparison to all other clusters. Interestingly, SC and cancer pathways including Pi3K and P53 were enriched in cluster 3 while cluster 4 was enriched with TNF pathways and multiple annotations for cancer (Fig. S2) . Additional analysis revealed a list of differentially expressed genes expressed by clusters 3 or 4 ( Fig. S3a-b) . Expectedly, genes related to hemi-desmosomes increased in clusters of basal cells (clusters 1, 3-4, 6-7), and desmosome-related and gap junction associated genes increased in clustered supra-basal cells (clusters 2, 5, 8-9) (Fig. S3c) . Interestingly, high levels of the extracellular matrix (ECM) components Lamb3 and Crim1 (Cysteine-Rich Motor Neuron 1), that regulates BMP and inhibits cell proliferation 57 , were found in the outer limbus cluster 3, whereas Col17a1 was higher in the inner limbus cluster 4. Taken together, this data illuminates the genetic signature of diverse cell states across the corneal epithelial lineage, and in particular, reveals the co-existence of two discrete limbal sub-compartments consisting of two distinct limbal epithelial cell populations. To further uncover the nature of outer and inner limbal epithelial cells, functional enrichment analysis (WebGestalt with ORA and KEGG) shed light on differences between clusters 3 and 4. Interestingly, among the top 10 functional categories, 4 identifiers involved cell proliferation while 3 involved cell movement (Fig. S4a) , suggesting that a central difference between the limbal clusters 3 and 4 lies in their cell cycle and motility characteristics. We thus aimed to gain insights regarding proliferation, clonal dynamics, centripetal movement, and discover the potential hierarchy between all clusters and within each limbal zone. For that purpose, we designed a quantitative genetic lineage tracing experiment for tracking all cell populations including potentially rare LSC populations, in a clonal and inducible manner. We established double transgenic UBC-Cre ERT2 ; Brainbow 2.1 mice that facilitated the tamoxifen inducible activation of Cre recombinase in all cell types in an efficient and unbiased manner under the control of the ubiquitin C (UBC) promoter (Fig. S4b ). In the absence of tamoxifen, Cre is expressed at the cytoplasm of all cell types of this mouse, thus unable to access the DNA. However, upon transient exposure to tamoxifen, Cre recombinase translocates into the nucleus where it can randomly induce DNA excision and/or inversion of the loxP sites in the brainbow 2.1 cassette, thus, allowing stochastic and irreversible expression of one out of four "confetti" fluorescent proteins (namely, nuclear green (GFP), cytoplasmic red (RFP), cytoplasmic yellow (YFP) or membrane cyan (CFP) fluorescent protein) (Fig. S4c) 58 . Two-month-old mice were treated with tamoxifen for 3-4 consecutive days to induce efficient "Confetti" labeling in all cell populations. For quantitative analysis of the limbal sub-compartments, wholemount staining of Gpha2 was performed to identify the outer limbus (647nm to avoid overlap with 4 Confetti colors and DAPI) while the contiguous 100m Gpha2-negative limbus was considered as inner limbus for calculations ( Fig. S4d) . Interestingly, at all time-points, the distribution of clones size (i.e., the number of basal cells in a clone) in the outer limbal epithelial zone was significantly lower compared to those found in the inner limbus and the peripheral cornea. Moreover, the average clonal size in the inner limbus was smaller than the size of clones in the corneal periphery ( Fig. 2a-d, e) . Two months post induction, a pattern of radial stripes emerged already in clones that held a "foot" in the inner limbus (Fig. 2b) . Four-months post induction, when the pattern of fully developed, radial stripes appeared. Most stripes extended from the inner limbus and only a few confetti stripes were sourced from the outer limbus (Fig. 2d, j) . The detection of stripes starting from both outer and inner limbal zones after very long-term tracing, suggests that both limbal zones contain LSCs. The uniform expression pattern of markers and clonal growth dynamics implied that the outer limbal epithelial cells might follow stochastic rules and neutral drift 59 . A main prediction of this approach is that the size of the clone would grow linearly with time 60 .In this case, the slope of this linear trend is proportional to the product of doubling rate and the probability of symmetric division in the epithelial plane (Methods). Our results suggest that clone size dynamics in all regimes can be captured by a linear growth and that the slopes are significantly different and ordered centripetally-each regime has a slope which is about twice higher than its neighboring zone ( Fig. 2f-g) . Another prediction of the neutral drift model, is that the clonal size distributions scale with their averages. That is, the size distribution of the clone size normalized by their mean should be similar. Fig. 2h shows the probability of having a clone that is larger than some value (one minus the cumulative probability distribution) for the normalized size distributions. All regimes show clear scaling of their clone size for all time points. Furthermore, the neutral drift model predicts that the number of clones should decrease inversely with the clone size (Fig. 2i ). In our case, there is a clear difference between the three different zones tested. While in the periphery the expected inverse relationship between size and number is similar to that of neutral drift, in the outer and inner regimes after four months the reduction in the number of clones is smaller than expected by that model. Taken together, this data suggests that the outer limbus and inner limbus contain a population of SCs that have similar stochastic dynamics but different doubling times. If the outer LSCs are the most primitive undifferentiated cells positioned at the tip of the cellular hierarchy, the inner limbus may host slightly committed LSCs. To further assess the hierarchical relations between clusters in the corneal epithelial lineage, we plotted the clusters using linear dimensional reduction using the Seurat algorithm for limbal/corneal epithelial clusters (clusters [3] [4] [5] [6] [7] [8] [9] . This algorithm infers the sequential changes of gene expression between clusters and thereby provides insights for the dynamics of biological processes, in this case cell differentiation. In line with lineage tracing data, this analysis suggested that cluster 3 (outer limbal basal cells) gives rise to cluster 4 (inner limbal basal cells) that derive clusters 6-7 (corneal basal cells) and clusters 8-9 (corneal wing and superficial cells), while cluster 5 cells (limbal superficial) that are indeed very similar to corneal superficial cells were positioned last (Fig. 2k) . Taken together, the combined quantitative lineage tracing and in silico analysis illuminates the dynamics and hierarchy between corneal epithelial cell populations. To rigorously test LSC growth dynamics by an alternative and direct approach, we next examined the cycling properties of outer and inner basal limbal epithelial cells by staining of Ki67 which labels cells that are at different stages of cell division (late G1, S, G2, M). In agreement, only 5.7±3.0% of outer limbal basal cells were Ki67 + , while 16.5±3.9% of inner limbal basal cells and 19.8±4.0% of corneal peripheral basal cells were Ki67 + ( Fig. 3a-b ). To corroborate this data, we performed 5-Ethynyl-2´-deoxyuridine (EdU) nucleotide analogue incorporation assay to identify undergoing DNA replication (S-phase). As illustrated in Fig. 3c , 2-month-old mice were intra-peritoneal injected with EdU, six hours later mice were sacrificed and cells that were in the S-phase of cell division during the interval were identified as EdU + cells by whole mount staining. Expectedly, EdU incorporation by outer limbal cells was much less frequent, as compared to that of basal epithelial cells at the inner limbus or peripheral cornea (Fig. 3d) . To examine the ability of outer limbal cells to enter active cell cycle in response to corneal injury, we performed large (2.5mm) corneal epithelial debridement following EdU injection. Evidently, 24-hours post injury, numerous outer (and inner) limbal cells entered cell cycle and were EdU positive. Taken together, this set of experiments suggests that the outer LSCs are slow-cycling cells, but they can quickly enter cell cycle and participate in corneal healing. To further examine the presence of slow cycling cells in the outer limbus, we performed an EdU pulse-chase experiment. As described in Fig. 3e , mice were injected with EdU twice a day for 15days ("pulse") followed by a "chase" period of 30-days at the absence of EdU. During the chase period the label declines by 50% within every cell division and the label becomes undetectable after few successive division rounds (~4), depending on assay type and sensitivity 61 38 . As expected, EdU + label retaining cells were predominantly found in the outer limbus, whereas rare cells were occasionally found in the inner limbus (Fig. 3f) , suggesting that during the 30-day chase period most quiescent outer LSCs exceeded sufficient division rounds and lost labeling and only few which underwent fewer replications could be detected. To validate that label retaining cells were epithelial cells, we performed co-staining of EdU + label retaining cells with Krt17 antibody (Fig. S1h) . The relative homogeneity in the expression of Gpha2/Ifitm3/Cd63 by the outer LSCs, together with the absence of Ki67 associated with broad slow growth of clones in the outer limbus, implied that the vast majority, perhaps entire, basal cell population in the outer limbal zone divides infrequently. To gain direct evidence on cell cycle of each population and evaluate the rate of cell division in each zone we performed a double nucleotide-analogue injection. As depicted in Fig. 3g, 1 .5 hours following injection of iododeoxyuridine (IdU, red), EdU (green) was injected, and 30 minutes later tissues were harvested, 8 stained and analyzed (Fig. 3h-i) . As detailed in Methods, S-phase length was first calculated as the ratio between the number of cells that were IdU + /EdU + (cells that were engaged with, but have not completed, DNA replication) and the number of cells that were IdU + /EdU -(cells that were in S-phase after IdU injection but have completed DNA replication already before EdU injection), multiplied by the interval length (1.5 hours). The estimated division frequency was calculated as the ratio between the total number of cells divided by the number of cells in S-phase, multiplied by the S-phase length 62 . This analysis suggests that outer limbal basal cells divide on average every ~8 days, inner limbal basal cells divide every ~4 days while corneal peripheral basal cells divide every ~3.5 days (Fig. 3i) . Taken together, this data strongly suggests that the outer limbus contains a predominant equipotent Krt15 + /Gpha2 + /Ifitm3 + /Cd63 + cell population of qLSCs while the inner limbus is occupied by a widespreas equipotent aLSCs. Cellular quiescence is linked with reduced biochemical activity including transcription activity 16 17 . Indeed, the cells in cluster 3 (qLSCs) displayed the lowest value of RNA reads per cell (Fig. 3j) . Previously, BMP was linked with epithelial SC quiescence while WNT repressed BMP pathway and induced SC activation 63 64 65 66 . In agreement, cells of cluster 3 (qLSCs) preferentially expressed the BMP-regulated transcription factors Id1/Id3 and the Wnt inhibitor, Sfrp1 67 , while Wnt3a was higher in cluster 4 of aLSCs (Fig. 3k) . Interestingly, cluster 3 expressed Cyclin D2 68 69 and the tumor suppressor Trp53 70 , both of which were linked with growth arrest induced by various stimuli. The relevance of qLSC to cell therapy is significant, therefore, we next explored the usefulness of the newly identified markers of murine qLSCs to human. Immunofluorescent staining revealed that KRT15, IFITM3 and GPHA2 are expressed by basal limbal epithelial cells (Fig. 4a) . Similar to staining of murine epithelium ( Fig. 1-3) , the labeling pattern of qLSC markers was not sporadic but rather wide, marking a large population of basal limbal epithelial cells. Interestingly, however, GPHA2 expression was drastically down regulated to hardly detectable levels upon cultivation of LSCs. Repression of GPHA2 (esiGPHA2) resulted in total inhibition of GPHA2 expression (Fig 4b) . This implies that GPHA2 expression depends on niche specific signal that was absent in vitro, and that GPHA2 is tightly associated with a slow cycling feature that is not maintained in vitro. Nevertheless, KRT15 and IFITM3 were highly expressed by undifferentiated limbal epithelial cells in vivo and in vitro (Fig. 4c ). In addition, calcium induced differentiation enhanced KRT12 and reduced KRT15, IFITM3 and P63 (Fig. 4c) . In a previous report, IFITM3 was linked with localization to endo-membranes where it protected embryonic SCs from viral infection 71 . In line, IFITM3 signal was confined to cellular vesicle structures in the cytoplasm of cultivated limbal cells (Fig. 4d) . To test whether IFITM3 influences stemness and differentiation, we performed a knock down experiment. To achieve efficient and specific knockdown, we used endoribonuclease-prepared silencing RNA (esiRNA) where enzymatic digestion of long double stranded RNAs produces a mixture of multiple short fragments of silencing sequences, thereby, enhancing specificity and efficiency. Efficient knock down of IFITM3 was evident by RNA and protein levels ( Fig. 4e-f) . Moreover, repression of IFITM3 resulted in a significant reduction in KRT15, and an increase in the expression of the differentiation associated corneal keratin KRT12 (Fig. 4e-f ). Taken together, this set of experiments suggests that IFITM3 positively controls the undifferentiated state of human qLSCs, and that the newly identified qLSC markers (IFITM3, GPHA2) can be used to identify human qLSCs. T cells serve as niche cells for qLSC regulating cell proliferation and wound closure 9 The striking segregation of qLSCs to the outer limbus suggests that this cell state must be regulated by the defined local microenvironment (niche). Interestingly, the outer limbus was clearly demarcated by wellstructured blood/lymph vasculature (Fig 5a) . Thin well-organized blood vessels typically penetrated the marginal inner limbus but never reached the corneal periphery. Lymph vessels were characteristically present in the outer but not the inner limbus. Additionally, the limbus was highly populated with Cd45 + immune cells, amongst them, many Cd45 + that displayed dendritic cell morphology, both in the stroma and in the epithelium (Fig. S5a) . Interestingly, we have identified a population of T cells in the outer limbus, many of which expressed the regulatory T cell markers, Cd4 and Foxp3 (Fig. 5b, S5a) and others expressed the cytotoxic T cell marker Cd8 (Fig. S5b) , suggesting that qLSCs may be regulated by T cells. Next, we explored the limbus of two laboratory mouse strains that fail to maturate T and B lymphocytes, namely, severe combined immunodeficiency (SCID) and non-obese diabetic SCID (NOD/SCID), and BALB/c mice served as a control group. Similar to C57BL/6 genetic background (Fig 1-3) , the outer limbus of BALB/c mice was hallmarked by Gpha2 + /Cd63 + /Ifitm3 + that infrequently expressed Ki67 (Fig. 5b-d, S5a -c, S5e). Curiously, an extensive reduction of Gpha2 and Cd63 proteins to levels that became barely detectable was found in SCID mice (Fig. 5b-c) as well as in NOD/SCID mice (Fig. S5c) , while other qLSC markers (e.g. Ifitm3) were unaffected (Fig. S5b) . Moreover, higher index of Ki67 labeling was found in the outer limbus (as well as inner limbus) of SCID mice (Fig. 5c-d) that displayed mild epithelial thickening (Fig. S5d) . In agreement, similar defects were recapitulated in NOD/SCID mice (Fig. S5c) . Since no B cells (Cd19 + ) were detected in the limbus (Fig. S5b) , we assumed that this phenotype is caused by the absence of T cells in SCID mice. To further address the involvement of T cells in qLSC regulation, we explored athymic Nude-Foxn1nu mice that lack T cell development. Here too, qLSC markers Gpha2 and Cd63 were dramatically attenuated while Ifitm3 was maintained, strengthening the involvement of T cells in qLSC regulation (Fig. S5e) . Finally, we aimed to substantiate the crosstalk of T cells -qLSC in adulthood, and exclude the possibility of developmental failure. To this end, we repressed the ocular immune system by topically applying the corticosteroid Dexamethasone (Fig. S5f) , or alternatively and more specifically, performed inhibition of regulatory T cells by sub-conjunctival injection of anti-Cd25 antibody (PC61.5). Intriguingly, 6-days post antibody injection, Cd63 and Gpha2 drastically decreased whereas cell proliferation increased (Fig. 5e-f ), strongly suggesting that T cells directly regulate quiescence in the outer limbus. Similar results were obtained following Dexamethasone treatment (Fig. S5f) . Finally, to link between T cell regulation and qLSC functionality, we performed corneal epithelial debridement and followed epithelial closure by fluorescein dye penetration. As shown in Fig. 5g -h, immunodeficient mice displayed delayed wound closure. Taken together, this set of experiments suggests that T cells, most likely Cd4+/Foxp3+ regulatory T cells, serve as a niche for qLSCs and play a critical role in quiescence maintenance, control of epithelial thickness and wound healing. In 1989, Cotsarelis and Lavker reported the identification of rare slow cycling cells in the limbus 10 . Since then, over the last 3-decades, much attention has been given for identifying such rare cells, however, no reliable marker has been found yet. Here we captured the signature of primitive qLSCs and propose that these cells are more prevalent than previously estimated. We showed evidence that qLSCs, which populate the basal layer of the outer limbus, quite uniformly express Krt15 + /Gpha2 + /Ifitm3 + /Cd63 + . The ~8 days estimated cell cycle length of qLSC fits well with quantitative lineage tracing, double nucleotide incorporation experiments, and nucleotide pulse-chase label retention assays. Indeed, the limbal/corneal epithelium becomes uniformly labeled following 2-weeks of nucleotide pulse 38 10 , i.e., suggesting that the length of cell cycle of both limbal and corneal epithelial cells is shorter than 14 days. Additionally, if assuming ~8 days for qLSCs, the interval of 30-days of chase in the absence of EdU (Fig. 3e-f ) should theoretically allow ~4 division rounds for outer qLSCs, and it is therefore not surprising that most outer limbal basal cells lost labeling while only few retained low EdU signal. In other words, this data fits well with abundant qLSCs model. Likewise, this model is supported by the quantitative analysis of Ki67 staining and EdU incorporation after short pulse (Fig. 3a-d) as in both cases, the outer limbus showed a widespread negativity for labeling. Our working model that posits abundancy of equipotent qLSCs fits well with a stochastic SC model that was proposed to describe SC dynamics in the epidermis, esophagus and gut epithelium 23 58 72 . In these tissues, abundant equipotent SCs that were shown to be extremely fast cycling, are at the tip of the cellular hierarchy. Epidermal and esophageal SCs divide every 2-3 days while gut epithelial SCs divide once a day, much faster as compared to qLSCs that divide every 8 days. This suggests that under distinct physiological conditions, tissues display tailored regenerative strategies for reasons of cost and benefit tradeoff. The two new LSC populations were well-segregated into well-defined limbal sub-compartments. This conclusion is supported by the observations that the outer and inner limbal cells differentially express specific markers and engaged pathways (Fig. 1, S2-4) , display distinct clonal growth and proliferation dynamics (Fig. 2-3) , and niche components (Fig 5) . The lineage tracing experiments implies that aLSCs play a key role in cornea replenishment under homeostasis, as most stripes emerged from this region. By contrast, a key feature of qLSCs is their ability to rapidly exit dormant state and enter the cell cycle in response to injury 73 (Fig. 3) . The cell cycle length of qLSCs is only approximately twice longer than that of aLSCs that were estimated to divide every ~4 days. However, this frequency of cell division is higher compared to some estimations for other tissue-specific slow cycling SCs 74 75 . The reason for such a drastic difference in SC cycling frequencies and its importance is poorly understood. In this sense, one could not exclude the possibility that an even slower cell population co-exists in the outer limbus, however, if it does, these cells must be very rare. Moreover, sub-fractionation of cluster 3 of qLSCs does not provide statistically relevant subclusters, again, suggesting the cells in the cluster were relatively highly homogenous. The analysis of the clone size distributions showed that they scale with the average number of clones in all the regions. This suggests that the underlying stochastic dynamics of all the regions are similar. The clone size dynamics are consistent with the neutral drift dynamics model and suggest that the doubling time in the inner limbus is twice the doubling time in the outer limbal regime, which is consistent with the cell cycle length measurements. In all regimes, the number of clones is going down with time, as expected. However, analyzing the quantitative relation between clone size and clone number reveals a striking difference between the outer and inner limbal regimes and the periphery. While all zones exhibit clone size distributions that are consistent with the neutral drift model, the decrease in the inner and outer regimes is slower than expected by this model. This saturation in clone number decrease rate might indicate that the clonal neutral competition is limited in these areas. This could be due to the tight spatial boundary conditions of the limbal regimes. Our working model (Fig. 6 ) is in agreement with the hematopoietic, hair follicle 18 , neural 20 and muscle 19 lineages, all of which incorporate qSCs that can produce aSCs 18 . What are the benefits and costs of having two distinct SC states? What is the mechanism that controls the transition between SC states? Is this model universal or otherwise only applies for particular tissues? Which physiological constraints influence SC division rates? Interestingly, epithelial SCs of the gut proliferate once a day 76 , while epidermal 77 and esophageal 78 SCs divide every ~3 days. In this context, it is tempting to hypothesize that the corneal transparency and location in the front of the eye makes it an extremely hazardous environment for SCs that must be well guarded especially because they must constantly proliferate to replenish corneal center with new cells. This must entail fundamental considerations for SC well-being such as localizing the SCs at the most well-protected zone (outer limbus) and accrediting SCs with infrequent division. The quiescence state of SCs is believed to be accompanied by reduction in DNA replication, metabolism, gene transcription and protein translation 16 . Notably, these processes were associated with molecular damage due to the production of toxic agents or errors in biosynthesis of macromolecules 79 80 . In line, attenuation of these activities by SC dormancy was proposed to delay SC aging 17 . Moreover, the reduced cumulating number of cell replications by quiescent SCs, may significantly attenuate the accumulation of mutations. Surprisingly, however, qSCs employ an error prone mechanism of non-homologous end joining to repair their DNA 81 82 . By contrary, frequently dividing SCs engage a non-error prone mechanism of homologous recombination for repairing DNA breaks 83 , and frequent division rate was shown to be required for the activation of repair pathways 84 . In line, frequently dividing SCs were identified in many tissues including the gut 22 , epidermis 85 ,and esophagus 24 . These cumulating findings revised the old dogma that considered quiescence as an integral feature of SCs. It also suggested that epithelial SCs are not quiescent and that the high turnover of such tissues is associated by a single pool of active SCs. The present study refines this conclusion and highlight the possibility that qSCs play an important role in corneal epithelial lineage and potentially in other epithelial tissues that may benefit from qSC reservoir. The regulation of quiescence is only partly understood, however, it clearly involves interactions with niche extracellular matrix proteins, secreted factors, and niche resident cells. At the molecular level, cell cycle inhibitors (e.g. P21, P27 and P57), tumor suppressor genes (e.g. retinoblastoma and P53) and soluble factors (e.g. BMP4, Dkk-1, sFRP, Wif and Shh) were found to positively regulate quiescence, while cytokines and Wnt activators (e.g. noggin) were associated with the exit from quiescence state and the transition into active SC state 18 16 . The bioinformatic analysis of limbal clusters suggests that qLSC engage Pi3K pathway, the BMP-regulated transcription factors Id1/3 63 64 65 , the tumor suppressor gene Trp53 16 , Cyclin D2 68 69 and the Wnt inhibitor Srf1 86 , all of which were linked with growth arrest. Of interest, aLSCs were enriched for TNF, WNT and Ap-1 pathways (Fig. S2-3 ) that were linked with proliferation or SC activation. Interestingly, Atf3 which is a structural protein of Ap-1 complex was shown to repress Id1 and enhance epithelial cell proliferation 87 . Future studies will address the mechanism by which these genes are regulated by outer/inner limbus specific ECM factors (Fig. S3) , metabolites or vasculature related factors, or influenced by niche cells (e.g. T cells). Likewise, future experiments will be needed to illuminate the mechanism that controls the quiescence and the transition into active state. The maintenance of LSC specific states, and the transition between them is most likely driven by niche cells (e.g. T cells, keratocytes, vasculature), ECM and/or biomechanical forces. It is therefore not surprising that quiescence (as well as Gpha2 expression) is lost ex vivo, whereas culture conditions are sub-optimal and lacking essential components of the niche. The co-localization of T cells with qLSC strongly implies that these cells may secret cytokines to control LSC quiescence, in line with previous studies where immune cells were identified as niche cells for epithelial SCs 88 . In the clinics, corticosteroids such as dexamethasone are being widely used to prevent uncontrolled corneal inflammation and neo-vascularisation, as these agents are considered as highly effective drugs for nonspecific suppression of inflammation. However, it has been shown that anti-inflammatory steroids must be used for short time as they may have adverse effects including corneal wound healing delay 89 . The awareness and recognition of such adverse effect is of particular importance given the observed interference of T cells -qLSC crosstalk. Future studies will be needed to illuminate these interactions and translate this knowledge into optimized treatments. Moreover, this knowledge may facilitate optimal qLSC cultivation for curing blindness in patients that suffer from "LSC deficiency" 90 . We identified a set of markers for detecting qLSCs that may allow identifying and perhaps purifying qLSCs. What is the functional role of these proteins? Do they control quiescence and/or maintain stemness? What controls their specific expression by qLSCs? Gpha2 (Glycoprotein hormone alpha-2) was found here to be uniquely expressed by qLSCs but its role is still unclear. Transgenic overexpression of Gpha2 in mice showed no gross phenotype 91 . Interestingly, we report a drastic reduction of Gpha2 levels in cases where the niche was disturbed, for example, when LSCs were disconnected from the niche and grown in vitro, or following immune cell repression or absence in vivo. In both cases, quiescence was lost, suggesting that Gpha2 may be regulated by T cells and controls quiescent cell state. By contrast, the interferon-induced transmembrane protein 3 (Ifitm3) was still expressed in the absence of immune cells in vivo and in vitro, suggesting that it is regulated by other niche means, not via a T cell-induced pathway, and that it does not control cell proliferation. Interestingly, both Ifitm3 and Cd63 reside in cellular endo membranes 92 . Moreover, Ifitm3 is known for its antiviral specific functions 93 , providing protection against entry and replication of several types of viruses including the COVID-19 virus 71 . Ifitm3 was shown to be highly expressed in embryonic, neural and mesenchymal SCs, suggesting that it may confer anti-viral activity to qLSCs. Interestingly, it was shown that Ifitm3 was involved in B cell related leukemia 94 and control embryonic germ cells 95 . Such evidence hints that ifitm3 may provide protection and control important functions of SCs. No major developmental disorder was reported regarding ifitm3 knockout mice, although the cornea or LSCs have not been investigated in that study 96 . Emb (Embigin), an adhesion molecule that marked qLSCs (Fig. S3a) , was shown to regulate hematopoietic SC quiescence and localization 97 . Furthermore, Emb was linked with quiescence of hair follicle SCs 55 while Gpha2 and Ifitm3 were shown to be expressed in epidermal SCs in the G0 state 98 . Future studies will be needed to illuminate the role of these genes in SC homeostasis and pathology, and to test whether their manipulation can contribute to advancing LSC therapy. In summary, this report provides a useful atlas that uncovers the main corneal epithelial cell populations, capturing the signature and the niche of quiescent and activated LSC states. These data open new research avenues for studying the mechanisms of cell proliferation and differentiation as well as the applications of LSCs in regenerative medicine. Ten eyes of 2.5 months old Krt15-GFP mice were enucleated, the limbus (with marginal conjunctiva and peripheral cornea) was dissected (~0.5 mm), tissues were pooled and incubated in trypsin (X10 Biological Industries) for 10 min 37°C 300 μl. Supernatant was collected into 10 ml (RPMI (Biological Industries) 10% chelated fetal calf serum). Trypsinization was repeated for 10 cycles adding fresh trypsin in each cycle. Cell suspension was centrifuged (8 min at 300g), re-suspended and filtered using Cell strainer (Danyel Biotech) to achieve 1200 cells/μL. An RNA library was produced according to the 10x Genomics protocol (Chromium Single Cell 3' Library & Gel Bead Kit v2, PN-120237) using 18000 input cells. Single cell separation was performed using the Chromium Single Cell A Chip Kit (PN-120236) . The RNAseq data was generated on Illumina NextSeq500, 150 bp paired-end reads, high-output mode (Illumina, FC-404-2005) according to 10X recommendations read 1-26 bp and read 2-98bp. Cell ranger (version 2.2.0) was used for primary analysis. Transcripts were mapped to the mm10 reference genome with the addition of the eGFP gene sequence. Downstream analyses including clustering, classifying single cells and differential expression analyses were performed using R package -Seurat (version 2.3.4) 46 . Low-quality cells (>5% mitochondrial UMI counts, or <200 or >8000 expressed genes 99 ) and genes (detected in <3 cells) were excluded, and eventually 14,941 genes across 4,738 cells were analyzed. In silico analysis of non-epithelial cell markers confirmed the absence (>0.1%) of corneal endothelial cells (Clrn1, Pvrl3, Gpc4, Slc9a7, Slc4a4, Grip1, Hrt1d, Irx2) , corneal stromal keratocytes (Vim, Thy1, S1004a4), goblet cells (Muc2, Muc5ac, Muc5b) , melanocytes (Dct, Mitf, Ptgs, Tyrp1, Melana, Tyr), T cells (Cd3d, Cd3e, Cd4, G2mb), B cells (Cd19, Cd79a, Cd79b). The only non-epithelial cells identified in our dataset were six cells that expressed MHC II (H2-Ab1 + ), 4 of them were Cd11c + /Cd207 + or Cd11c + /Cd39 + dendriticlike cells, 2 others were Cd45 + /Adgre + or Cd45 + /Cd64 + macrophage-like. Analysis of pan epithelial genes (expression of Krt14, Krt12, Cdh1, Epcam and Cldn7) confirmed that most (>99%) cells were ocular epithelial cells. For pathway analysis, a list of differentially expressed genes between chosen clusters was analyzed using WebGestalt 56 . Over-Representation Enrichment Analysis (ORA) was used for pathway analysis with KEGG functional database. Animals, tissue preparation and staining: Animal care was according to the ARVO Statement for the Use of Animals in Ophthalmic and Vision Research. R26R-Brainbow 2.1 (#013731), K14-Cre ERT2 (#005107), Krt15-GFP (#005244) and UBC-CreERT2 (#008085) were from Jackson Laboratories (Bar Harbor, ME). To induce Cre recombinase activity, 4 mg/day Tamoxifen (T5648, Sigma, St. Louis, MO) dissolved in corn oil was intraperitonealy injected (200 μl) for 3-4 consecutive days, as previously reported 29 30 . NOD/SCID, SCID, Nude, and BALB/c were from Envigo RMS Ltd (Israel). Dexamethasone (Sigma) was dissolved in PBS (0.5%) and administered as drops (15 µl) on ocular surface of anesthetized (2% Isoflurane) mice, for 10 minutes every 12 hours for 4 consecutive days. For wounding, mice were anesthetized (2% Isoflurane) and injected intramuscular with analgesics Buprenorphine (0.03 mg/ml, 50 µl). Central corneal wounding (2 mm diameter) was performed using an ophthalmic rotating burr (Algerbrush) under fluorescent binocular. The wounded corneas were stained (1% fluorescein), and the wounded area was imaged and quantified (NIS-Elements analysis D). For CD4 + FoxP3 + regulatory T cell depletion, subconjunctival injection of 30 µl (250 µg) anti-Cd25 antibody (Clone:PC61, antiCd25 IL-2Rα) or rat IgG (control) (BioCell InVivoMAb) 100 was performed. Single intraperitoneal injection of 200 μl (7.5 mg/ml) EdU (Sigma) was performed and 6 hours later tissues were processed. For pulse-chase, EdU was injected every 12 hours for 14 consecutive days and tissues processed after additional 30-days. Isolated corneas were fixed (2% PFA, 1 hour) and stained (Click-iT, Invitrogen) according to the manufacturer's instructions followed by whole mount staining protocol (described above) for other markers staining. For the calculation of length cycle (Tc) 1.7 mg/ml IdU (Abcam) was injected and 1.5 hours later, 7.5 mg/ml EdU was injected and tissues harvested 30 minutes later. The calculation of length cycle (Tc) was calculated according to formula described by 62 Staining, imaging and quantifications: For paraffin sections, eyes were fixed (overnight, 4% formaldehyde), washed briefly with phosphate buffered saline (PBS), incubated in increasing ethanol, Xylene and paraffin (60°C) and then embedded in paraffin blocks, as previously described 33 . Sections (5 μm) were prepared, antigen retrival performed with unmasking solution (Vector Laboratories, H3300), blocked (0.2% Tween20 and 0.2% gelatin, 2 hours), incubated with primary antibody (overnight, 4°C), secondary antibodies (Invitrogen, 1:500, 2 hours) followed by 4′,6-diamidino-2-phenylindole (DAPI) staining, and mounting (Thermo Scientific). For H&E (Sigma) staining, sections were rehydrated, incubated with Hematoxylin (7 minutes), washed (tap water, 15 minutes), stained with Eosin (3 minutes), dehydrated, incubated with Xylene (15 minutes) and mounted (Thermo Scientific). For frozen sections, eyes were fixed (4 hours, 4% formaldehyde), washed briefly (PBS), incubated in 30% sucrose (overnight) and embedded in optimal cutting temperature compound (OCT). Sections (5 μm) were fixed (4% paraformaldehyde), permeabilized (0.1%, TritonX-100, 20 minutes), blocked (1% BSA, 1% gelatin,2.5% Normal goat serum, 2.5% Normal donkey serum, 0.3% TritonX-100, 1 hour), incubated with primary antibody (overnight, 4°C), secondary antibodies (1:500, 1 hour), DAPI and mounting (Thermo Scientific). For whole mount, corneas were isolated, fixed (2% formaldehyde, 2 hours, room temperature), permeabilized (0.5% Triton, 5 hours), blocked (0.1% TritonX-100, 2% Normal donkey serum, 2.5% BSA, 1 hour), primary antibody (overnight, 4°C on a shaker), secondary antibodies (1:500, 1 hour), followed by DAPI, tissue flattening under a dissecting binocular and mounting (Thermo Scientific), as previously described 33 Images were acquired using 20X/0.8 M27 Plan-Apochromat objective on LSM 880 airyscan laser scanning confocal system (Zeiss, Oberkochen, Germany) or by Nikon Eclipse NI-E upright Microscope using Plan Apo λ 20x objective. For computerized quantification analyses we used Nis-elements Analysis D software, the outer limbus was marked by Gpha2 or Cd63 staining, the adjacent 100 µm (or entire inner limbus if K15-GFP used) of the inner limbus was calculated, while 100 µm peripheral cornea was quantified and 5 different field were calculated from each cornea (n=3-5). In vitro experiments: Human limbal rings from cadaveric corneas were obtained under the approval of the local ethical committee and declaration of Helsinki. Cells cultivation, differentiation, transfection, Western blot and real time polymerase chain reaction (qPCR) analyses were performed as previously described 101 . We used esiRNA against EGFP (EHUEGFP, Sigma), IFITM3 (EHU222801, Sigma) or GPHA2 (EHU036941, Sigma) and cells were collected 5-6 days after transfection. Primers for qPCR were: GAPDH (GCCAAGGTCATCCATGACAAC, CTCCACCACCCTGTTGCTGTA), IFITM3 (CTGGGCTTCATAGCATTCGCCT, AGATGTTCAGGCACTTGGCGGT), KRT15 (GACGGAGATCACAGACCTGAG, CTCCAGCCGTGTCTTTATGTC), KRT12 (TGAATGGTGAGGTGGTCTCA, TTTCAGAAGGGCAAAAAGGA). Clone size and number analysis: As described previously 27 23 60 , one of the main predictions of the neutral drift dynamics model is that the average number of cells in a given clone as a function of time is given by, where  is the fraction of basal cells that divide, 2r is the probability for symmetric division and  is the division rate. The number of clones is expected to decrease with time as, . Thus, the relation between the clone size and clone number is given by Liner fit of the clone size was performed using Matlab linear regression model. Estimating the clone size cumulative distribution was done using empirical CDF estimation using Matlab. Statistical analysis: Data are presented as means +/-SD. T-test, ANOVA (analysis of variance) followed by Bonferroni test, or Kolmogorov-Smirnov test were performed using GraphPad Prism software, as indicated in legends, to calculate p-values. Differences were considered to be statistically significant from a p-value below 0.05. Statistical significance was calculated using t-test (*, p-value<0.05). Abbreviations: Cj., Conjunctiva; Peri., Periphery. Figure 6 : Working hypothesis. Schematic illustration of our working model, depicting the anatomy of the limbal sub-compartments and LSC populations. We propose that the outer limbus is occupied by widespread qLSCs that play an important role as a SC reservoir for wound repair, barrier maintenance and corneal replenishment. The inner limbus contains abundant aLSCs that are actively engaged in corneal replenishment. The maintenance and differentiation of LSC populations is regulated by niche components including blood and lymph vasculature, stromal cells, and particularly T cells that control qLSC proliferation. (H) EdU pulse/chase experiment described in Fig. 3e-f that was performed to detect slow cycling cells. The square zone on the right shows enlarged view of the dashed region on the left. Arrow shows EdU + /Krt17 + limbal epithelial cells at the outer limbus. Scale bars are 50 μm. Abbreviations: Conj., Conjunctiva; Peri., Periphery. Figure S2 : Prediction of pathways enriched in clusters 3 or 4. In silico analysis performed to predict pathways that were enriched in cluster 3 or 4 (in comparison to all other clusters) using the Webgestalt algorithm. Figure S3 : Differential expression of selected families of genes. Heatmap presentation shows the average expression of the indicated genes across clusters. Genes enriched in cluster 3 or cluster 4 are shown in A and B, respectively, and selected family-related genes are shown in (C). Outer limbus Krt15-GFPregion Krt15-GFP + region 106 ± 81 µm 60 ± 8 µm 179 ± 61 µm Defining Adult Stem Cells by Function, not by Phenotype Epithelial stem cells: turning over new leaves Advances in stem cell research and therapeutic development Somatic stem cell heterogeneity: Diversity in the blood, skin and intestinal stem cell compartments Defining Adult Stem Cell Function at Its Simplest: The Ability to Replace Lost Cells through Mitosis The biology of hematopoietic stem cells Hematopoiesis: an evolving paradigm for stem cell biology Disparate differentiation in mouse hemopoietic colonies derived from paired progenitors Stem cells: units of development, units of regeneration, and units in evolution Existence of slow-cycling limbal epithelial basal cells that can be preferentially stimulated to proliferate: Implications on epithelial stem cells In vivo proliferation and cell cycle kinetics of long-term self-renewing hematopoietic stem cells In Vivo Fate Analysis Reveals the Multipotent and Self-Renewal Capacities of Sox2+ Neural Stem Cells in the Adult Hippocampus Label-retaining cells reside in the bulge area of pilosebaceous unit: implications for follicular stem cells, hair cycle, and skin carcinogenesis The intestinal epithelial stem cell: the mucosal governor Asymmetric division and cosegregation of template DNA strands in adult muscle satellite cells Mechanisms, Hallmarks, and Implications of Stem Cell Quiescence Good and Bad of Stem Cell Aging Coexistence of quiescent and active adult stem cells in mammals Tissue-specific stem cells: lessons from the skeletal muscle satellite cell Concise review: Quiescent and active states of endogenous adult neural stem cells: identification and characterization The tortoise and the hair: slow-cycling cells in the stem cell race Identification of stem cells in small intestine and colon by marker gene Lgr5 A single type of progenitor cell maintains normal epidermis Cellular heterogeneity in the mouse esophagus implicates the presence of a nonquiescent epithelial stem cell population Clonal analysis of stem cells in differentiation and disease Live imaging of stem cells: answering old questions and raising new ones Kinetics of cell division in epidermal maintenance Corneal epithelial stem cells: deficiency and regulation Lineage tracing of stem and progenitor cells of the murine corneal epithelium A Method for Lineage Tracing of Corneal Cells Using Multi-color Fluorescent Reporter Mice Tracing the fate of limbal epithelial progenitor cells in the murine cornea Lineage tracing in the adult mouse corneal epithelium supports the limbal epithelial stem cell hypothesis with intermittent periods of stem cell quiescence Corneal-Committed Cells Restore the Stem Cell Pool and Tissue Boundary following Injury Peripheral (not central) corneal epithelia contribute to the closure of an annular debridement injury Limbal epithelial stem cells: role of the niche microenvironment Limbal epithelial crypts: a novel anatomical structure and a putative limbal stem cell niche Assessment of corneal substrate biomechanics and its effect on epithelial stem cell maintenance and differentiation Characterization of slow cycling corneal limbal epithelial cells identifies putative stem cell markers Comparison of limbal and peripheral human corneal epithelium in tissue culture Location and clonal analysis of stem cells and their differentiated progeny in the human ocular surface Cytokeratin 15 can be used to identify the limbal phenotype in normal and diseased ocular surfaces C/EBPdelta regulates cell cycle and self-renewal of human limbal stem cells ABCB5 is a limbal stem cell gene required for corneal development and repair ABCG2 transporter identifies a population of clonogenic human limbal epithelial cells p63 identifies keratinocyte stem cells Integrating single-cell transcriptomic data across different conditions, technologies, and species Structure, development and function of cytoskeletal elements in nonneuronal cells of the human eye Patterns of cytokeratin and vimentin expression in the human eye Corneal expression of SLURP-1 by age, sex, genetic strain, and ocular surface health Corneal integrins and their functions Cell cycle protein expression and proliferative status in human corneal cells Tight junction transmembrane protein claudin subtype expression and distribution in human corneal and conjunctival epithelium Hierarchical expression of desmosomal cadherins during stratified epithelial morphogenesis in the mouse Organization and formation of the tight junction system in human epidermis and cultured keratinocytes Capturing and profiling adult hair follicle stem cells WebGestalt: an integrated system for exploring gene sets in various biological contexts A novel role for CRIM1 in the corneal response to UV and pterygium development Intestinal crypt homeostasis results from neutral competition between symmetrically dividing Lgr5 stem cells Universal patterns of stem cell fate in cycling adult tissues Quantitative Clonal Analysis and Single-Cell Transcriptomics Reveal Division Kinetics, Hierarchy, and Fate of Oral Epithelial Progenitor Cells Computer simulation of neutral drift among limbal epithelial stem cells of mosaic mice Foxg1 is required for specification of ventral telencephalon and region-specific regulation of dorsal telencephalic precursor proliferation and apoptosis Signaling through BMPR-IA regulates quiescence and long-term activity of neural stem cells in the adult hippocampus BMP signaling and its pSMAD1/5 target genes differentially regulate hair follicle stem cell lineages The inhibitor of differentiation isoform Id1b, generated by alternative splicing, maintains cell quiescence and confers self-renewal and cancer stem cell-like properties Wnt signaling maintains the hair-inducing activity of the dermal papilla Axin2 marks quiescent hair follicle bulge stem cells that are maintained by autocrine Wnt/β-catenin signaling Increased expression of cyclin D2 during multiple states of growth arrest in primary and established cells Glucose regulates cyclin D2 expression in quiescent and replicating pancreatic β-cells through glycolysis and calcium channels p53 regulates hematopoietic stem cell quiescence Intrinsic Immunity Shapes Viral Resistance of Stem Cells Long-lived keratin 15+ esophageal progenitor cells contribute to homeostasis and regeneration Injury Activates Transient Olfactory Stem Cell States with Diverse Lineage Capacities Hematopoietic stem cells reversibly switch from dormancy to self-renewal during homeostasis and repair Quantitative proliferation dynamics and random chromosome segregation of hair follicle stem cells The adult Drosophila posterior midgut is maintained by pluripotent stem cells Defining the cellular lineage hierarchy in the interfollicular epidermis of adult skin Esophageal stem cells--a review of their identification and characterization The DNA-damage response in human biology and disease DNA Damage in Stem Cells Hematopoietic stem cell quiescence promotes error-prone DNA repair and mutagenesis Bcl-2 and accelerated DNA repair mediates resistance of hair follicle bulge stem cells to DNA-damage-induced cell death Molecular regulation of stem cell quiescence DNA repair mechanisms in dividing and non-dividing cells Epidermal homeostasis: a balancing act of stem cells in the skin Defining the epithelial stem cell niche in skin A self-enabling TGFbeta response coupled to stress signaling: Smad engages stress response factor ATF3 for Id1 repression in epithelial cells Two to Tango: Dialog between Immunity and Stem Cells in Health and Disease Corticosteroids and corneal epithelial wound healing Eyes on the prize: limbal stem cells and corneal restoration A glycoprotein hormone expressed in corticotrophs exhibits unique binding properties on thyroid-stimulating hormone receptor Antiviral Role of IFITM Proteins in African Swine Fever Virus Infection The broad-spectrum antiviral functions of IFIT and IFITM proteins Monoclonal antibody to the interferoninducible protein Leu-13 triggers aggregation and inhibits proliferation of leukemic B cells The generation of spermatogonial stem cells and spermatogonia in mammals Normal germ line establishment in mice carrying a deletion of the Ifitm/Fragilis gene family cluster Embigin Regulates HSPC Homing and Quiescence and Acts As a Cell Surface Marker for a Niche Factor-Enriched Subset of Osteolineage Cells Defining the Design Principles of Skin Epidermis Postnatal Growth Classification of low quality cells from single-cell RNA-seq data In vivo depletion of CD4+FOXP3+ Treg cells by the PC61 anti-CD25 monoclonal antibody is mediated by FcgammaRIII+ phagocytes SOX2 Regulates P63 and Stem/Progenitor Cell State in the Corneal Epithelium Outer InnerPeri.