Abstract In vivo conversion of nonneuronal cells into neurons is a proposed strategy to replace neurons lost to CNS injury or disease. Glia-to-neuron trans-differentiation by viral vector–mediated GFAP mini-promoter–driven NeuroD1 remains hotly debated. Developing inducible, lineage-traceable transgenic mice, we find that astrocyte-to-neuron conversion is restricted to a specific time window within the lesion core of injured spinal cord and brain. Spatiotemporal lineage-mapping combined with single-cell transcriptomics reveals that ectopic NeuroD1 induces astrocyte-to-neuron conversion specifically in lesion cores via transit-amplifying OLIG2^+ progenitors during early injury phase, but not in late phases or in nonreactive astrocytes. Neither a loss-of-function NeuroD1 mutant nor stemness-reprogramming factor SOX2 induces astrocyte-to-neuron conversion. However, contrary to previous reports, the neuronal-like cells generated by NeuroD1 lack mature neuroelectrical properties, limiting their functional integration into neural circuits. Together, our findings establish a spatiotemporal framework for NeuroD1-driven glia-to-neuron conversion, revealing a mechanistic shift from direct astrocyte conversion toward transit-amplifying intermediates and highlighting the functional immaturity of NeuroD1-converted neurons. __________________________________________________________________ NeuroD1 converts astrocytes to neurons via transit-amplifying progenitors, but the resulting neurons lack functional maturity. INTRODUCTION Most neurodegenerative diseases of the central nervous system (CNS) are characterized by the progressive loss of neurons. The adult mammalian brain and spinal cord have limited regenerative capacity ([46]1–[47]3). Although diverse therapeutic strategies for CNS injuries have been investigated, including stimulation of endogenous neural stem cell (NSC) differentiation, administration of neuroprotective compounds, and transplantation of exogenous NSCs or neurons ([48]4–[49]8), efficacies in replacing lost neurons and generating functional neuronal activity and networks are limited. The direct in situ conversion of endogenous nonneuronal cells into functional neurons through fate conversion in the brain and spinal cord following injury has been proposed as a potential therapeutic approach ([50]9–[51]12). Neurons, astrocytes, and oligodendrocytes, the three fundamental neural cell types in the CNS, share the same NSC or progenitor cell origins during development ([52]13). Astrocytes are highly abundant and heterogeneous in the CNS and can be transient amplified in response to injury, representing a potential source to replace neuronal loss after injury or disease ([53]13–[54]16). It has been reported that neurogenic factors, which are active in radial glial stem cells during embryonic neurogenesis but down-regulated during differentiation, can be reactivated in postnatal and adult astrocytes to promote neuronal formation ([55]10, [56]17). Diseases and injuries can trigger astrocyte remodeling and reactivity including non-proliferative or proliferative responses ([57]18, [58]19). The plasticity of reactive astrocytes in response to injury positions them as a potential candidate for induced neurogenesis or lineage conversion ([59]20–[60]22). NeuroD1, a member of the basic helix-loop-helix transcription factor family, regulates neuronal differentiation during embryonic development ([61]23) and is vital for the survival and maturation of adult-born neurons ([62]24, [63]25) through reprogramming of chromatin and transcriptional landscapes ([64]26, [65]27). Previous studies reported that adeno-associated virus (AAV) vector–mediated expression of NeuroD1 induces glial fibrillary acidic protein (GFAP)–expressing astrocytes to trans-differentiate into mature neurons ([66]28, [67]29), proposing a therapeutic potential of NeuroD1 in animal models of spinal cord injuries, stroke, and epilepsy as well as other neurological disorders ([68]30–[69]36). However, recent studies have raised questions about the conclusions of these studies as to authentic conversion of astroglia into neurons ([70]37–[71]40). Mis- or leaky expression of the neurogenic factor and reporter in already existing neurons may occur when using AAVs carrying a GFAP mini-promoter to drive NeuroD1 expression ([72]37–[73]40). Thus, reporters used in this context do not faithfully indicate the glial origin of labeled neurons. Furthermore, the permissive state of astrocytes for NeuroD1-induced trans-differentiation remains controversial ([74]17, [75]41). While some reports suggest that NeuroD1 can drive neuronal conversion even in nonreactive astrocytes ([76]42), this remains a subject of debate. Additionally, the specific time window during which NeuroD1 induces neuronal formation from astrocytes following CNS injury remains poorly understood. Now, in vivo tools with transgenic animal models for cell lineage– and temporal-specific NeuroD1 induction are lacking, hindering robust in vivo genetic fate mapping needed to precisely define the cell state and temporal dynamics of astroglia-to-neuron conversion. To address the potential issues of mis-expression and leakage of reporters in viral vectors in endogenous neurons, we developed an inducible, lineage-traceable transgenic knock-in mouse line for NeuroD1 expression and used these mice to examine resident astrocyte-to-neuron conversion in different regions following spinal cord or brain injury. We found that ectopic NeuroD1 expression in reactive astrocytes in the lesion core via a transit-amplifying intermediate, but not in distal injury region, drove the formation of neuronal cells in the injured spinal cord and brain. Trans-differentiation is observed when NeuroD1 expression is induced during the early post-injury phase but not at later stages. However, contrary to the previous reports, we find that the newly converted neuron-like cells induced by NeuroD1 did not exhibit the electrophysiological characteristics of typical mature neurons. Thus, our robust, reproducible genetic models resolve key controversies by demonstrating NeuroD1 induction of authentic astrocyte-to-neuron conversion following CNS injury. In addition, our studies highlight that reprogramming through transit-amplifying intermediates, rather than direct astrocyte conversion, represents a unified mechanism for in vivo neuronal conversion while uncovering that the NeuroD1-converted neurons fail to achieve functional characteristics of mature neurons. These findings underscore the need to develop therapies that promote fully functional neurons in CNS diseases and injuries. RESULTS Generation of inducible and lineage-traceable NeuroD1 knock-in mice To determine the effect of NeuroD1 on astrocyte-to-neuron conversion after CNS injury, we generated an inducible transgenic knock-in line for NeuroD1 expression. The Rosa26 locus was targeted with a transgene containing a loxP-flanked STOP cassette, a mouse Neurod1 cDNA, and a P2A–green fluorescent protein (GFP) reporter. The NeuroD1-P2A^GFP fusion allows simultaneous expression of NeuroD1 and GFP, with the GFP signal used to trace NeuroD1-expressing cells ([77]Fig. 1A). To activate transgene expression in astrocytes, we cross-bred a well-validated, astrocyte-specific Aldh1l1-CreERT2 line ([78]43) with a Rosa-tdTomato (Ai9) reporter line (control mice) ([79]44) or with the Rosa-NeuroD1-P2A^GFP transgenics, hereafter referred to as iNeuroD1^GFP mice ([80]Fig. 1A). Fig. 1. Ectopic NeuroD1 expression results in reactive astrocyte-to-neuronal cell conversion in the lesion core after spinal cord injury. [81]Fig. 1. [82]Open in a new tab (A) Schematic of Rosa-CAG-loxP-STOP-loxP-tdTomato reporter mice, activated with Aldh1ll-CreERT2 to generate tdTomato-expressing astrocytes in control mice (left). Schematic of CAG-loxP-STOP-loxP-iNeuroD1^GFP transgene inserted into the Rosa26 locus, activated with Aldh1ll-CreERT2 to generate NeuroD1-GFP expression in astrocytes in iNeuroD1^GFP mice (right). (B) Experimental scheme showing spinal cord injury (SCI) on day 0, following by TAM injection (dpi 5 to 8), in control and iNeuroD1^GFP mice at dpi 10. (C) Representative images showing the co-labeling of GFP (green) and NeuroD1 (red). Scale bars, 25 μm. (D) Western blot analysis for NeuroD1 in iNeuroD1^GFP and control mice with or without injury, normalized to β-actin. (E) Representative image of GFAP-labeled spinal cord showing the lesion core and distal region after SCI in control mice. Blue, 4′,6-diamidino-2-phenylindole (DAPI) stain. Scale bar, 250 μm. (F) Immunostaining for NeuN and tdTomato (tdT) or NeuroD1^GFP in the lesion core of control and iNeuroD1^GFP mice, respectively, at dpi 10. Scale bars, 50 μm. Right: Quantification of the percentage of reporter-positive cells expressing NeuN in the lesion core. n = 3 mice per group; means ± SEM. (G) Representative immunolabeling images for NeuN and tdTomato (tdT) or GFP in the distal region of control and iNeuroD1^GFP mice, respectively, at dpi 10. Scale bars, 50 μm. Right: Quantification of the percentage of reporter positive cells also positive for NeuN in the distal region. n = 3 mice per group. n.d., not detectable. To validate astrocyte-specificity of exogenous gene expression in control and iNeuroD1^GFP mice, spinal cord tissue was harvested after 3 days of tamoxifen (TAM) administration (100 mg/kg/day) to induce gene expression (fig. S1A) ([83]45). After post-induction at day 5, tdTomato and GFP signals were predominantly detected in GFAP^+ astrocytes in both control and iNeuroD1^GFP mice, and these signals were not detected in neurons (NeuN^+), oligodendrocyte progenitors (PDGFRα^+), or microglia (Iba1^+) (fig. S1, B and C). These data indicate that ectopic NeuroD1 expression does not convert normal astrocytes to NeuN^+ neurons in the uninjured spinal cord To next investigate the fate of ectopic NeuroD1-expressing cells after injury, control and iNeuroD1^GFP mice were subjected to dorsal hemisection spinal cord injury ([84]46). TAM was administered from 5 to 8 days post-injury (dpi) to induce NeuroD1 or tdTomato (control) expression ([85]Fig. 1B) ([86]45). GFP and NeuroD1 were co-expressed in the spinal cords of iNeuroD1^GFP mice, either with or without an injury, at 2 days after TAM administration to activate transgene expression in Aldh1ll^+ astrocytes ([87]Fig. 1C). GFP expression thus serves as a surrogate readout of TAM-inducible NeuroD1 expression under both physiological and pathological conditions. The induction of NeuroD1 expression by TAM-treatment was confirmed by Western blot analysis of spinal cord tissues in iNeuroD1^GFP mice at 10 days post-induction ([88]Fig. 1D). Notably, spinal cord injury alone did not induce endogenous NeuroD1 expression in the spinal cord of control mice ([89]Fig. 1D). Together, these data indicate that TAM induces ectopic NeuroD1 expression in the Aldh1ll^+ astrocytic lineage in iNeuroD1^GFP mice. Ectopic NeuroD1 expression at early injury phase induces astrocyte-to-neuronal cell conversion To investigate the effect of NeuroD1 expression on the fate of Aldh1l1-CreERT2–expressing astrocytes, we performed spinal cord injury in control and iNeuroD1^GFP adult mice at postnatal day 60 (P60) and administered TAM (100 mg/kg/day) ([90]45) from dpi 5 to 8. During this post-injury period, reactive gliosis with proliferative GFAP^+ astrocytes peaks at dpi 5 (fig. S2, A and B), compared to the acute phase at dpi 3 or the late phase at dpi 10, suggesting a transient amplifying phase of reactive astrocytes during the early phase of injury around dpi 5. Because astrocyte limitans borders around lesions extend ~0.5 mm from the injury site after spinal cord injury ([91]47, [92]48), we divided injured spinal cord tissues into the lesion core and peri-injury distal regions ([93]Fig. 1E), which are enriched with proliferative and non-proliferative reactive astrocytes, respectively (fig. S2, A and B). TAM administration induced expression of tdTomato in control mice and NeuroD1/GFP in iNeuroD1^GFP mice ([94]Fig. 1F). At dpi 10, most of GFP-expressing cells in the lesion core of iNeuroD1^GFP mice were NeuN^+ neuronal cells (77.6 ± 1.9%), whereas, in control mice, none of the tdTomato^+ astrocytes were co-labeled with NeuN ([95]Fig. 1F). These data suggest that ectopic NeuroD1 induces robust NeuN^+ neuronal cell formation from reactive astrocytes in the lesion core. In contrast, in the distal injury regions of control and iNeuroD1^GFP mice, none of the reporter-labeled cells expressed NeuN ([96]Fig. 1G). This indicates that, in non-proliferative reactive astrocytes at the distal injury region, NeuroD1 expression did not convert astrocytes into NeuN^+ neuronal cells. Thus, during the transient amplifying period of reactive astrocytes after spinal cord injury, NeuroD1 can only induce the robust expression of NeuN in neuron-like cells in reactive astrocytes in the lesion core and not in distal regions. NeuroD1 ectopic expression before injury promote astrocyte-to-neuron conversion An injury could potentially activate the Aldh1l1 promoter in preexisting neuronal cells, a confound of the AAV-mediated delivery approach ([97]49). To ensure that NeuroD1 was only expressed in preexisting resident astrocytes directed by Aldh1l1-CreERT2 in adult mice, we administered TAM to iNeuroD1^GFP mice for 3 days before injury to induce Cre expression in resident astrocytes and then performed spinal cord injury ([98]Fig. 2A). Immunostaining for the astrocytic markers GFAP and S100β confirmed the presence of GFP reporter–labeled astrocytes within the lesion core ([99]Fig. 2B). At dpi 21, ~54.4% of NeuroD1^GFP+ astrocytes became NeuN^+ neuronal-like cells in the lesion core, whereas no NeuroD1^GFP+/NeuN^+ cells were observed in the distal injury region ([100]Fig. 2, C and D). The NeuN induction rate compared favorably with the 77.6% of NeuN^+GFP^+ cells detected in iNeuroD1^GFP mice when the injury preceded the TAM-induction ([101]Fig. 1F). These data suggest that the NeuN^+ neuronal-like cells formed post-injury are derived from preexisting astrocytes with the forced NeuroD1 expression before the injury as opposed to ectopic transgene expression in endogenous neurons. Fig. 2. NeuroD1-induced astrocyte-to-neuron conversion in a time- and cell-state–dependent manner after spinal cord injury. [102]Fig. 2. [103]Open in a new tab (A) Pre-injury experimental scheme to induce NeuroD1^GFP expression in Aldh1l1^+ astrocytes, followed by SCI and analysis at dpi 21. (B) Representative immunolabeling images of S100β and GFAP in the lesion core at 21 days after SCI. Scale bar, 50 μm. (C) Representative immunolabeling images of NeuroD1^GFP and NeuN in the lesion core and at distal regions in iNeuroD1^GFP mice at dpi 21. Scale bars, 50 μm. (D) Quantification of the percentage of NeuN^+ NeuroD1^GFP+ cells. n = 3 mice per group; means ± SEM. (E) Post-injury experimental scheme to induce SCI (day 0), followed by daily BrdU injection from days 2 to 5, TAM from days 5 to 8 to induce NeuroD1^GFP expression in Aldh1l1^+ astrocytes, and analysis at dpi 10. (F) Representative immunolabeling images of NeuroD1^GFP, BrdU, and NeuN in the lesion core at dpi 10. Scale bars, 25 μm and (zoomed-in region) 10 μm. (G) Quantification of the percentage of BrdU^+NeuroD1^GFP+NeuN^+ cells n = 3 mice per group; means ± SEM. (H) Post-acute experimental scheme to induce NeuroD1^GFP expression in Aldh1l1^+ astrocytes, followed by SCI and analysis at dpi 18. (I) Representative immunolabeling images of S100β and GFAP in the lesion core at dpi 18. Scale bar, 50 μm. (J) Immunostaining for NeuN and NeuroD1^GFP in the lesion core of iNeuroD1^GFP mice at dpi 18. Scale bar, 50 μm. (K) Quantification of the percentage of NeuroD1^GFP+cells expressing NeuN in the lesion core. n = 3 mice per group; means ± SEM. (L) Quantification of the percentage of Ki67^+NeuroD1^GFP+cells in the lesion core. n = 3 mice per group; means ± SEM. n.d., not detectable. Conversion is restricted to the injury core via transient-amplifying intermediates After injury, astrocytes in proliferative and non-proliferative states are detected in the lesion core and peri-injury distal region, respectively ([104]18). To trace the fate of proliferative reactive astrocyte subpopulations in iNeuroD1^GFP mice, we administrated 5-bromo-2′-deoxyuridine (BrdU) from dpi 2 to 5, the peak period of reactive astrocyte self-amplification ([105]48), to label transit-amplifying cells and followed by TAM administration from dpi 5 to 8 to induce NeuroD1 expression ([106]Fig. 2E). We detected BrdU^+/NeuroD1^GFP+/NeuN^+ intermediates in the lesion core but not in distal regions ([107]Fig. 2F), and most of BrdU-labeled NeuroD1^GFP+ cells expressed NeuN (91.6 ± 4.4%, [108]Fig. 2G). Thus, transient-amplifying intermediates with NeuroD1 overexpression are converted into NeuN-expressing neuronal-like cells in the lesion core after spinal cord injury, whereas non-proliferative astrocytes in the peri-injury distal region are not. NeuroD1 does not induce astrocyte-to-neuron conversion at late injury phase To determine whether the temporal window of reactive gliosis affects NeuroD1-induced neuronal conversion, we delayed TAM administration to induce NeuroD1 expression at dpi 14 to 16 ([109]Fig. 2H), corresponding to a late phase of gliosis when astrocytes exhibit reduced proliferative activity (fig. S2, A and B). At dpi 18, in the lesion core of iNeuroD1^GFP mice, where reactive astrocytes marked by GFAP and S100β were abundant ([110]Fig. 2I), we did not detect GFP-expressing cells co-labeled with NeuN ([111]Fig. 2, J and K) or Ki67^+ proliferative astrocytes ([112]Fig. 2L). Compared to neuronal conversion by NeuroD1 induction in the injury core during the early post-injury phase (dpi 5 to 8), the failure of neuronal conversion at later stages suggests that NeuroD1-induced astrocyte-to-neuron conversion is spatiotemporally restricted, occurring primarily during the transit-amplifying phase of reactive astrocytes within the lesion core. Immature neuroblasts are formed during astrocyte-to-neuron conversion To track the transition of reactive astrocytes into neurons, we analyzed iNeuroD1^GFP mice that had been subjected to spinal cord injury and treated with TAM from dpi 5 to 8 ([113]Fig. 3A). At the early phase of injury, most of NeuroD1^GFP+/NeuN^+ cells (~98% at dpi 10 and 78% at dpi 30) co-expressed GFAP and exhibited nonneuronal or glia-like morphology, with bipolar and multipolar branches ([114]Fig. 3, B and C). In contrast, at 17 weeks post-injury (wpi), only a small population of NeuroD1^GFP+/NeuN^+ cells (13.5 ± 5.6%) continued to express GFAP. These data suggest that, between dpi 10 and 30, NeuroD1^GFP+/NeuN^+ cells are at an intermediate GFAP^+ stage of their differentiation into more neuronal-like cells. Fig. 3. Reactive astrocytes undergo NeuroD1-induced conversion through intermediate states to neuron-like cells. [115]Fig. 3. [116]Open in a new tab (A) Experimental plan to identify intermediate states of gene expression after injury. SCI was induced on day 0, followed by TAM-induction of NeuroD1^GFP expression in Aldh1l1^+ astrocytes at day 7, and analysis at dpi 10 and 14 and at wpi 17. (B) Representative immunolabeling images of NeuroD1^GFP, GFAP, and NeuN in the lesion core at dpi 10 and 30 and at wpi 17. Scale bars, 50 μm and (zoomed-in region) 10 μm. (C) Quantification of the percentage of GFAP^+NeuN^+NeuroD1^GFP+ (upper graph) or GFAP^−/NeuN^+NeuroD1^GFP+ (lower graph) cells. n = 3 mice per group; means ± SEM. ***P < 0.001, one-way analysis of variance (ANOVA) followed by Holm-Šídák’s multiple comparisons test. (D) Representative immunolabeling images of NeuroD1^GFP or the tdTomato reporter with DCX in the lesion core of iNeuroD1^GFP and control mice, respectively, at dpi 10, 14, and 24. Scale bars, 50 μm and (zoomed-in region) 10 μm. (E) Quantification of number of DCX^+ cells in lesion field of spinal cord injury. n = 3 mice per group; means ± SEM. n.d., not detectable. (F) Quantification of the percentage of DCX^+ reporter^+ cells. n = 3 mice per group; means ± SEM. n.d., not detectable. (G) Representative immunolabeling images of NeuroD1^GFP with OLIG2 and NeuN in the lesion cores from three animals at dpi 14. Scale bars, 50 μm and (zoomed-in region) 10 μm. (H) Summary schematic showing that NeuroD1-induced conversion of reactive astrocytes to neuron-like cells occurs only in the lesion core and passes through an intermediate state. To further characterize early-stage NeuN^+/NeuroD1^GFP+ cells, we analyzed expression of doublecortin (DCX), a marker of transient immature neuroblasts during early adult neurogenesis ([117]50). At dpi 10, ~36% of NeuroD1^GFP+ cells in the lesion core were DCX^+, whereas no DCX expression was detected in tdTomato^+ cells in control mice ([118]Fig. 3, D to F). Between dpi 14 and 24, the expression of DCX in NeuroD1^GFP+ cells gradually declined ([119]Fig. 3, D to F). Following spinal cord injury, a substantial proportion of reactive astrocytes within the lesion core at the early stages exhibited up-regulated expression of oligodendrocyte transcription factor 2 (OLIG2) (fig. S2, C and D), which was also detected in proliferative reactive astrocytes (fig. S2E) ([120]51). The OLIG2 transcription factor is essential for the proliferation of reactive astrocytes ([121]51) and serves as a marker for transit-amplifying progenitors with neurogenic potential in the adult brain ([122]52, [123]53). Notably, most of NeuroD1^GFP+ cells (~63.5%) expressed OLIG2 in the lesion core ([124]Fig. 3G). These results suggest that NeuroD1^GFP+/NeuN^+ cells may arise from an OLIG2^+ amplifying progenitor state and that NeuroD1-mediated astrocyte-to-neuron conversion occurs through immature neuroblast intermediates ([125]Fig. 3H). NeuroD1 induces reactive astrocyte-to-neuron conversion in the brain injury core To assess the effects of NeuroD1 in different CNS regions, we investigated the NeuroD1-mediated astrocyte fate switch in the brains of iNeuroD1^GFP mice by using two injury models: focal cerebral ischemia and stab wound injury. In the focal cerebral ischemia brain injury model, extensive GFAP^+ reactive astrocytes were induced surrounding the lesion, accompanied by cortical neuronal loss ([126]Fig. 4, A and B) ([127]54). When TAM was administered from dpi 5 to 8 to induce NeuroD1^GFP expression ([128]Fig. 4A), ~62.8% of NeuroD1^GFP+/GFAP^+ cortical astrocytes expressed the neuronal marker NeuN within the lesion core by dpi 10 ([129]Fig. 4, C and D). GFAP^+ was not co-expressed with NeuN in the distal region in the injured brain of iNeuroD1^GFP mice as well as both lesion core and distal regions of control mice ([130]Fig. 4D). Moreover, DCX was expressed by ~54.3% of NeuroD1^GFP+ cells, indicative of an immature neuronal state ([131]Fig. 4, E and F), while, among NeuroD1^GFP+/NeuN^+ cells, ~72.3% expressed OLIG2 ([132]Fig. 4, G and H). These data suggest that NeuroD1-mediated astrocyte-to-neuron conversion also occurs through immature intermediates in the focal cerebral ischemia injury model. Fig. 4. NeuroD1 converts reactive astrocytes to neuron-like cells in focal stroke and stab wound brain injury models. [133]Fig. 4. [134]Open in a new tab (A) Experimental scheme to induce focal stroke injury. (B) GFAP immunolabeling of lesion. Scale bar, 250 μm. (C) Representative images of GFAP, NeuroD1^GFP, and NeuN in the lesion core of iNeuroD1^GFP mice at dpi 10. Scale bar, 50 μm. (D) Quantification of the percentage of NeuN^+NeuroD1^GFP+ cells in the distal region and lesion core. n = 3 mice per group; means ± SEM. (E) Representative images of reporter and DCX in control and iNeuroD1^GFP mice at dpi 10. Scale bars, 50 μm and (zoomed-in region) 10 μm. (F) Quantification of number of DCX^+ cells (top) and percentage of reporter^+ DCX^+ cells (bottom) in lesion field. n = 3 mice per group; means ± SEM. (G) Representative images of NeuroD1^GFP and OLIG2 at dpi 10. Scale bars, 50 μm and (zoomed-in region) 10 μm. (H) Quantification of OLIG2^+ /NeuroD1^GFP+NeuN^+ cells. n = 3 mice per group; means ± SEM. (I) Experimental scheme of stab wound (SW) injury. (J) GFAP immunolabeling of lesion core. Scale bar, 100 μm. (K) Representative images of GFAP, NeuroD1^GFP, and NeuN in the lesion core of iNeuroD1^GFP mice. Scale bar, 50 μm. (L) Quantification of the percentage of NeuN^+NeuroD1^GFP+ cells in different region. n = 3 mice per group; means ± SEM. n.d., not detectable. (M) Representative images of NeuroD1^GFP and DCX in iNeuroD1^GFP mice. Scale bars, 50 μm and (zoomed-in region) 10 μm. (N) Quantification of number of DCX^+ cells (left) and percentage of NeuroD1^GFP+DCX^+ cells (right) in lesion. n = 3 mice per group; means ± SEM. (O) Representative images of NeuroD1^GFP and OLIG2 at dpi 10. Scale bars, 50 μm and (zoomed-in region) 10 μm. (P) Quantification of OLIG2^+ /NeuroD1^GFP+NeuN^+ cells. n = 3 mice per group; means ± SEM. The stab wound brain injury results in minimal neuron loss ([135]51, [136]55). In this model, we examined the NeuroD1 effects on glia-to-neuron conversion by administrating TAM from dpi 5 to 8 to induce NeuroD1^GFP expression ([137]Fig. 4, I and J). Of the NeuroD1^GFP+ cells in the lesion core, ~22.6% expressed NeuN ([138]Fig. 4, K and L). In contrast, GFP and NeuN did not co-localize in regions distal to the stab wound ([139]Fig. 4, K and L). Additionally, we also observed that ~9.3% NeuroD1^GFP+ cells expressed DCX ([140]Fig. 4, M and N) and ~72.2% of NeuroD1^GFP+/NeuN^+ cells expressed OLIG2 ([141]Fig. 4, O and P). These data from two additional models of brain injury demonstrate that ectopic expression of NeuroD1 is able to convert proliferating or progenitor-like reactive astrocytes in the lesion core into neuron-like cells in the injured brain. Single-cell transcriptomics reveals NeuroD1-mediated neuronal expansion within lesion core To comprehensively determine the cellular identities of NeuroD1^GFP-expressing cells following spinal cord injury, we performed single-nucleus RNA sequencing (snRNA-seq) of the lesion core and the peri-injury regions in iNeuroD1^GFP mice at dpi 10. Single cells dissociated from spinal cord segments including the lesion core and distal-injury regions from three mice were subjected to snRNA-seq using the 10X Genomics Chromium platform. We focused on the EGFP-expressing nuclei identified by the EGFP-WPRE pre-RNA transcripts after filtering out low-quality cells (see Materials and Methods; [142]Fig. 5A). Uniform Manifold Approximation and Projection (UMAP) revealed distinct differences in the transcriptomic features of EGFP-expressing nuclei isolated from the lesion core and those from the distal region ([143]Fig. 5, B and C). Using unbiased clustering, 176 and 288 EGFP-expressing nuclei from the lesion core and distal-injury site samples, respectively, were grouped into four clusters ([144]Fig. 5, B and C). Cell clusters were annotated using a harmonized atlas of known cell types in spinal cord ([145]56). Three neural cell types were represented: neurons, astrocytes, and oligodendrocyte lineage cells ([146]Fig. 5D). Analysis of cell proportions across these clusters revealed notable compositional differences between the lesion core and distal region, with a predominance of neuronal populations in the lesion core ([147]Fig. 5, B and E). Gene Ontology (GO) terms of differentially expressed genes overrepresented in the lesion core compared to distal regions included terms involved in nervous system development, neuron projection, and neurogenesis ([148]Fig. 5F). UMAP plots showed that EGFP-expressing cells with neuronal gene expression were substantially enriched in the population unique to the lesion core ([149]Fig. 5G). Thus, these single-cell profiling analyses demonstrate that ectopic expression of NeuroD1 in reactive astrocytes in the lesion core results in an increase in neuronal cell populations, consistent with the formation of NeuroD1^GFP+/NeuN^+ neuronal cells at the lesion core in iNeuroD1^GFP mice. Fig. 5. snRNA-seq data reveal NeuroD1-induced transcriptomic changes. [150]Fig. 5. [151]Open in a new tab (A) Illustration of tissue sampling and snRNA-seq experimental and sequence-based cell sorting workflow. (B) UMAP plots showing cells identified in the lesion core and distal region of iNeuroD1^GFP mice obtained at dpi 10. (C) UMAP visualization of 176 and 288 spinal cord nuclei sequenced from lesion core and peri-injury samples, respectively, color coded based on signature gene expression. (D) UMAP plots showing annotations of cells identified in lesion core and distal region. (E) Cell proportions across different clusters in the lesion core and distal region. (F) GO analysis of differentially expressed genes in the EGFP-WPRE sequence-sorted cells from lesion core compared with those from the distal region. (G) Gene expression patterns of neuronal genes and astrocyte genes notably changed in the lesion core relative to distal region. NeuroD1 loss-of-function mutant cannot induce astrocyte-to-neuron conversion To validate that NeuroD1 neurogenic activity is necessary for the conversion of astrocytes into neuron-like cells in our model, we generated mice that inducible-express a previously described NeuroD1 mutant that lacks the transactivation domain ([152]Fig. 6, A and B) ([153]57). Using an anti-NeuroD1 antibody targeting the N-terminal epitope for Western blot, we confirmed the expression of the truncated NeuroD1 protein in Aldh1l1-CreERT2; Rosa-NeuroD1-mut (iNeuroD1-mut^GFP) mice 2 days after TAM administration ([154]Fig. 6C). Following TAM treatment, brains were immunolabeled for GFAP and NeuroD1, which confirmed that the NeuroD1-mutant protein was ectopically expressed in GFAP^+ astrocytes in iNeuroD1-mut^GFP mice ([155]Fig. 6D). Fig. 6. Astrocyte-to-neuronal cell conversion requires the neurogenetic activity of NeuroD1. [156]Fig. 6. [157]Open in a new tab (A) Schematic of NeuroD1 and mutant NeuroD1(1 to 155 amino acids) [NeuroD1(1-55aa)]. AD1, activation domain 1; AD2, activation domain 2; bHLH, basic helix-loop-helix. (B) Diagram of CAG-loxP-STOP-loxP-NeuroD1-mutant ^GFP transgene targeted to the Rosa26 locus to generate iNeuroD1-mut^GFP mice. (C) Western blot analysis for NeuroD1 and NeuroD1 mutant protein in tissues harvested 2 days after last TAM injection. Nd1, iNeuroD1^GFP mice; Nd1Mut, NeuroD1-mut^GFP mice. (D) Representative images of brain tissue stained for GFP (green), GFAP (magenta), and NeuroD1 (red). Scale bars, 10 μm. (E) Representative immunolabeling images of NeuroD1-mut^GFP+ cells at lesion core for NeuN and GFAP at dpi 10 and 28. Scale bars, 50 μm. (F) Quantification of NeuroD1-mut^GFP+NeuN^+ cells in lesion core at dpi 10 and 28 (n = 3 mice per group; means ± SEM). n.d., not detectable. (G) Representative immunolabeling images for NeuroD1^GFP+NeuN^+ converted neuronal cells (arrows) and NeuroD1^GFP-NeuN^+ preexisting neurons (arrowheads) of iNeuroD1^GFP mice at dpi 10. Soma area was labeled by dashed circle. Scale bars, 50 μm and (zoomed-in region) 25 μm. (H) Box plot showing the range and median of somas for NeuroD1-induced converted neurons and preexisting neurons at dpi 10 and 30 and at wpi 17. n = 60 cells per group from 3 animals. ***P < 0.001, Student’s t test. We next used the same spinal cord injury model to determine the effects of NeuroD1-mut on the fate of reactive astrocytes. We did not detect any NeuN expression in GFAP^+ reactive astrocytes at the lesion core of iNeuroD1-mut^GFP mice at dpi 10 after TAM administration from dpi 5 to 8, whereas the iNeuroD1-mut^GFP+/GFAP^+ reactive astrocytes were abundant ([158]Fig. 6, E and F). To assess whether the NeuroD1 mutant has weak neurogenic activity rather than a complete loss of function, we examined iNeuroD1-mut^GFP mice at a later stage, at dpi 28 ([159]Fig. 6E). No NeuroD1-mut^GFP+/NeuN^+ neuron-like cells were detected ([160]Fig. 6F). These results indicate the crucial role of NeuroD1 neurogenic activity in the conversion of proliferating or dedifferentiated reactive astrocytes into neuron-like cells. Stemness reprogramming by SOX2 alone is ineffective in astrocyte-to-neuron conversion Expression of a stemness-reprogramming factor, SOX2, was reported to facilitate resident astrocytes into DCX^+ neuroblasts that subsequently mature into synapse-forming neurons in the injured adult spinal cord ([161]58). To investigate whether ectopic SOX2 expression could induce neuronal conversion from reactive astrocytes, we generated a TAM-inducible transgenic knock-in line for SOX2 expression carrying an mCherry reporter (fig. S3A). We cross-bred the inducible SOX2-expressing transgenic mice with Aldh1l1-CreERT2 mice to produce Aldh1l1-CreERT2; Rosa-Sox2 mice, hereafter called iSox2^mCherry mice, and validated that mCherry^+ cells co-expressed SOX2 in the spinal cord followed by TAM administration (fig. S3B). Next, we performed the same dorsal hemisection spinal cord injury (day 0) and injected TAM from dpi 5 to 8 in iSox2^mCherry mice (fig. S3C). In contrast to iNeuroD1^GFP mice, in the iSox2^mCherry mice, only a small fraction (less than 3%) of mCherry^+NeuN^+ cells were detectable in the lesion core region at dpi 10, with no mCherry^+ cells showing NeuN^+ expression at dpi 23 (fig. S3, D and E). Similarly, very few mCherry^+DCX^+ cells were observed in the lesion core region at dpi 11 (fig. S3F). Thus, compared to NeuroD1, ectopic expression of SOX2 does not effectively induce NeuN^+ neuronal cell trans-differentiation from proliferating or dedifferentiated reactive astrocytes in the mouse model after spinal cord injury. Neuronal cells resulting from NeuroD1 induction are not physiologically functional Even when using NeuroD1 with neurogenic activity, the induced neuronal cells displayed smaller soma sizes compared to preexisting neurons ([162]Fig. 6, G and H), suggesting a potential deficit in electrical activity and integration into neural circuits ([163]59). This morphological feature prompts us to investigate the electrophysiological characteristics of these induced neurons. To evaluate the electrophysiological properties of NeuroD1^GFP+/NeuN^+ cells, we performed whole-cell patch-clamp recordings of these cells in the lesion cores of acute ex vivo spinal cord slices and filled the recorded cells biocytin ([164]Fig. 7A). Together with avidin staining of biocytin-filled NeuroD1^GFP cells, immunostaining confirmed that they were positive to neuron marker NeuN ([165]Fig. 7B). Only GFP^+ cells positive to avidin and NeuN were analyzed. We evaluated the generation of action potentials at three time points: (i) wpi 3, the early conversion stage; (ii) wpi 9, the stage in previous studies when mature neurons were detected ([166]30); and (iii) wpi 17, a later stage. Fig. 7. NeuroD1-induced neuronal cells are not functionally mature. [167]Fig. 7. [168]Open in a new tab (A) Schema of patch-clamp recording in spinal cord slices, along with a differential interference contrast image of a NeuroD1-induced cell filled with avidin. Scale bar, 25 μm. (B) Representative immunolabeling images of recorded (iNeuroD1^GFP+/avidin^+) cells for NeuN. Scale bar, 25 μm. (C) Representative voltage responses to step current injections of NeuroD1^GFP+ NeuN^+ cells at wpi 3, 9, and 17. (D) Phase plots (top, voltage; bottom, slope) of NeuroD1^GFP+ NeuN^+ cells at wpi 3, 9, and 17. dv/dt, the rate of change of membrane voltage over time. (E) Representative repetitive action potentials in response to depolarizing current injection of a preexisting neuron. (F) Pie chart of the percentage of recorded NeuroD1^GFP+NeuN^+ neuron-like cells (n = 27 cells) with different firing patterns at wpi 3. (G) Representative whole-cell currents of recorded cells in the absence and presence of 1 μM tetrodotoxin (TTX) at wpi 3. Note the blockade of the inward Na^+ currents in the presence of TTX. (H) Representative immunolabeling images of NeuroD1^GFP+ cells with vGAT and Map2 at wpi 17. Scale bars, 10 μm. (I) Quantification of synaptic puncta in each NeuroD1-induced neuron-like cells and preexisting neurons. n = 3 mice per group; ≥60 cells were counted; means ± SEM. *P < 0.05, Student’s t test. To characterize the electrophysiological profiles, we generated phase plots of the waveforms to quantify the action potential depolarizing and repolarizing slope ([169]Fig. 7, C and D). At the early stage wpi 3, the NeuroD1^GFP+/NeuN^+ cells exhibited heterogeneous electrophysiological properties. Approximately 84% of the NeuroD1^GFP+cells retained glia-like membrane properties without action potential generation during stepwise current injections, while 12% of cells generated only single spikelet when subjected to strong current stimulation (>500 pA, 500 ms in duration), and 4% exhibited a smaller spikelet followed by sustained depolarization ([170]Fig. 7, C, D, and F). This is in contrast to the preexisting neurons that had repetitive action potentials evoked by depolarizing current steps in current-clamp mode ([171]Fig. 7E). These neuron-like cells exhibited a resting membrane potential ~−70 mV with no signs of instability or abnormal depolarization, suggesting that they are healthy cells. Consistently, sodium currents, which are essential for generating action potentials, were detected in the NeuroD1^GFP+/NeuN^+ cells and verified through tetrodotoxin blockade ([172]Fig. 7G). Given that sodium currents are absent in normal astrocytes ([173]60), the electrophysiological properties of the NeuroD1^GFP+/NeuN^+ cells resemble immature neurons. Similarly, at wpi 9, we failed to detect repetitive action potentials in NeuroD1^GFP+/NeuN^+ cells, and these cells exhibited a weak single spikelet ([174]Fig. 7, C and D). Even at the later stage wpi 17, NeuroD1^GFP+/NeuN^+ cells had mainly nonneuronal electrophysiology ([175]Fig. 7, C and D). Although a set of NeuroD1^GFP+/NeuN^+ cells generated a smaller spikelet, the depolarizing responses were substantially greater than that of typical neurons ([176]Fig. 7C). These observations indicate that the neuronal-like cells converted from astrocytes by expression of NeuroD1 do not have electrophysiological properties of mature neurons. Last, we further compared the circuit integration between NeuroD1-converted neuron-like cells and preexisting neurons. For the converted NeuroD1^GFP+ neuron-like cells in the lesion core, we observed that MAP2^+ cells with large somas were surrounded by synaptic markers vGAT (vesicular γ-aminobutyric acid transporter) and vGlut2 (vesicular glutamate transporter 2) ([177]Fig. 7H). However, the numbers of synapses were much lower in the converted neurons than in preexisting neurons ([178]Fig. 7, H and I). Together, although transient amplifying reactive astrocytes in the lesion core can be converted into neuronal-like cells, these converted neurons have limited synaptic connectivity and neuronal activity. DISCUSSION Neurodegenerative diseases and CNS injuries are challenging to treat due to the limitation of regenerative capacity of neurons in the adult brain and spinal cord ([179]1, [180]2). A potential therapeutic approach is direct in situ cellular reprogramming, in which endogenous nonneuronal cells, such as astrocytes or microglia, are converted into neurons in vivo ([181]61, [182]62). Astrocytes are the most common and abundant cells in the CNS, which makes them a potential cell source for reprogramming in neuronal replacement therapy ([183]14–[184]16, [185]63). Previous studies with viral vectors as delivery vehicles reported that ectopic expression of NeuroD1 can convert reactive astrocytes into mature and functional neurons under various CNS injury and disease conditions ([186]28–[187]35). However, recent studies have raised concerns regarding the authenticity of the astrocytic origins of converted neurons, particularly made using AAV or lentiviral vectors carrying a GFAP mini-promoter for the delivery of NeuroD1 ([188]37–[189]40). Evidence suggests that the human GFAP mini-promoter used in viral vectors exhibits mis-expression in preexisting neurons when used to drive the expression of neurogenic transcription factors like NeuroD1 ([190]39, [191]64, [192]65). This results in the expression of NeuroD1 and its reporter not only in astrocytes but also in endogenous neurons, confounding conclusions about the authentic neuronal conversion of resident astrocytes. In this study, we aimed to address these controversies by developing inducible, lineage-traceable transgenic mouse models to precisely evaluate astrocyte-to-neuron conversion in vivo. Our findings from multiple inducible transgenic mouse models provide compelling evidence that ectopic expression of NeuroD1 in resident reactive astrocytes following acute spinal cord or brain injury induces the formation of neuron-like cells specifically within the lesion core. However, in contrast to previous reports, our results also uncover previously unrecognized functional limitations in the converted neurons. Inducible genetic models demonstrate the authenticity of astroglia-to-neuron conversion The authenticity for the astroglia-to-neuron is hotly debated ([193]37–[194]40). Our inducible genetic animal models offer definitive spatiotemporal clarity and reproducibility that viral systems lack. With spatiotemporal cell lineage mapping at pre- and post-injury stages using the inducible transgenic knock-in mouse models, our findings resolve the controversies regarding NeuroD1-induced astrocyte-to-neuron conversion. Our data provide compelling evidence that the ectopic expression of NeuroD1 in astrocyte lineage cells after injury to the spinal cord or brain leads to the formation of neuron-like cells specifically in the lesion core. By integrating single-cell transcriptomics, loss-of-function mutants, and spinal cord and brain injury models, we provide molecular and cellular evidence to clarify the astrocytic identity of NeuroD1-mediated neuronal conversion. TAM administration may influence estrogen receptor–mediated gene expression, potentially affecting neural progenitor cell fate during spinal cord injury ([195]66, [196]67). To control for this, we used Ai9-tdT; Aldh1l1-CreER mice and administered TAM in the same manner as in NeuroD1; Aldh1l1-CreER mice. Notably, no NeuN expression was observed in these control animals following TAM treatment, indicating that TAM alone does not induce neuronal conversion from reporter-positive astrocytes ([197]Fig. 1F and fig. S1), unlike ectopic expression of NeuroD1. Thus, TAM treatment alone does not promote astrocyte-to-neuron conversion in the spinal cord injury model, although we cannot fully exclude the possibility that TAM affects estrogen receptor–mediated gene expression in ways that may influence NeuroD1-driven neuronal fate conversion. Astrocyte-to-neuron conversion is spatiotemporally restricted In contract to the previous report of the neuronal conversion from nonreactive astrocytes ([198]42), we found that NeuroD1-mediated neuronal conversion does not occur in physiologically normal, nonreactive astrocytes in the intact spinal cord and brain. In addition, ectopic expression of NeuroD1 is unable to trans-differentiate the non-proliferative reactive astrocytes into neurons in the peri-injury region. In contrast, within the lesion core, most of NeuroD1-expressing cells adopt neuron-like characteristics, suggesting that NeuroD1-mediated neuronal conversion from reactive astrocytes is region specific. Notably, this conversion does not occur beyond the early injury phase, underscoring a specific time window after injury for NeuroD1-induced astrocyte reprogramming. Thus, our findings suggest that the astrocyte-to-neuron conversion is restricted to a specific time window after injury and to the core region of the injured spinal cord and brain, providing previously unrecognized insights into the contextual dependencies of NeuroD1-mediated glia-to-neuron reprogramming. NeuroD1-mediated astrocyte-to-neuronal conversion is cell-state dependent Our findings demonstrate that NeuroD1 promotes neuronal trans-differentiation only from a distinct reactive astrocyte transit-amplifying state in the injury core, in contrast to the previously reported all states of astrocytes including nonreactive astrocytes ([199]42). We found that Aldh1l1^+ mature astrocytes can be dedifferentiated or transitioned into OLIG2-expressing transit-amplifying intermediates after acute injury. The NeuroD1-mediated fate conversion mainly targets these proliferative or transit-amplifying intermediates such as those expressing OLIG2. Notably, OLIG2 is critical for reactive astrocyte proliferation ([200]51) and is expressed in the transit-amplifying (type C) progenitors for neurogenesis in the adult brain ([201]52, [202]53). This suggests that transit-amplifying reactive astrocytes formed in the core after acute injury may impart reprogramming competence to the NeuroD1-mediated neuronal cell conversion process. In contrast, in regions distal to the injury or during the late phase of injury, where reactive astrocytes remain in a non-proliferative state, NeuroD1 expression fails to induce neuronal conversion. Thus, our findings reveal a previously unrecognized spatiotemporal and cell-state–specific conversion of resident reactive astrocytes to neurons by ectopic NeuroD1 post-injury in both spinal cord and brain. Transit-amplifying intermediates represent a unified mechanism for neuronal conversion Our spatiotemporally specific lineage analyses identify the intermediate states such as BrdU^+NeuN^+ or OLIG2^+NeuN^+ transitional cells, suggesting that the conversion of astrocytes into neuron-like cells is contingent on the transit-amplifying state of reactive astrocytes within the lesion core. Consistently, retroviral vectors carrying pro-neurogenic factors, which selectively transduce mitotic cells, are sufficient to promote glia-to-neuron conversion after injury ([203]33, [204]68), although these retroviruses may potentially target injury induced, dividing NSCs or microglia for neuronal reprogramming. In addition, alteration of cell extrinsic signaling such as Notch signaling can program astrocytes to a proliferative ground state to facilitate glia-to-neuronal fate conversion ([205]20–[206]22, [207]68). We showed that mature astrocytes transition to the transit-amplifying intermediate progenitor state in the injury core, rendering them permissive to NeuroD1-mediated neuronal trans-differentiation. Consistently, previous studies have shown that mature astrocytes proliferating after injury undergo transcriptional dedifferentiation and reprogramming into a transient progenitor-like state ([208]48) and that astrocytes isolated from injured adult cerebral cortex during this time period can give rise to neurospheres with pluripotent potential that are able to generate neurons in vitro, whereas mature astrocytes from adult uninjured cortex do not have this potential ([209]69, [210]70). These findings further support the conclusion that it is not surprising that the forced ectopic over-expression of a powerful fate-directing transcription factor like NeuroD1 in the dedifferentiated, proliferating, or progenitor-like reactive astrocytes in the immediate vicinity of injuries would drive them to differentiate into immature neuronal-like cells. Together, these observations suggest transit-amplifying intermediates as a convergent mechanism of cell-state plasticity underlying in vivo glia-to-neuron conversion. Thus, by shifting the mechanistic framework from direct conversion of astrocytes toward transit-amplifying intermediates, our study reveals a unified mechanism for in vivo reprogramming and provides a translational potential by leveraging transit-amplifying intermediates to enhance neuronal reprogramming for CNS repair. Stem-cell reprogramming alone is not sufficient to promote neuronal trans-differentiation To further validate the specificity and necessity of NeuroD1 neurogenic activity, we explored the effects of a transcription-inactive NeuroD1 mutant. We found that a transcription-inactive NeuroD1 mutant failed to induce astrocyte-to-neuron conversion, suggesting the specificity and necessity of NeuroD1 neurogenic activity in the process. This contrasts with AAV-GFAP mini-promoter–mediated mis-expression of a NeuroD1 mutant, which induces reporter expression in preexisting endogenous neurons ([211]39), further indicating the technical issues introduced with this mode of delivery. These observations underscore the importance of using functionally active NeuroD1 for the glia-to-neuron conversion process. Furthermore, to investigate the effects of the stem cell-reprogramming factor SOX2, previously reported to induce glia-to-neuron trans-differentiation with viral vectors ([212]58), we found that transgenic expression of SOX2 could not effectively convert reactive astrocytes into NeuN^+ neurons in the lesion core in our transgenic animal systems. These observations suggest that expression of a stemness-reprogramming factor SOX2 alone is insufficient to promote neuronal trans-differentiation, but it may facilitate or cooperate with other neurogenic factors to induce neurogenesis ([213]71). Neurons derived from NeuroD1-expressing astrocytes are not functionally mature Contrary to previous reports ([214]28–[215]36), our electrophysiological analyses unexpectedly revealed that neuronal-like cells derived from NeuroD1-induced astrocyte conversion lacked the electrophysiological properties of mature neurons. Notably, the soma size of NeuroD1-induced neurons derived from astrocytes was smaller than that of normal neurons ([216]Fig. 6, G and H). These data suggest that, while NeuroD1 can initiate the reprogramming of reactive astrocytes into neuron-like cells, these cells do not reach full maturity and functionality. Moreover, their integration into existing neural circuits was limited, as they formed fewer synaptic connections compared to preexisting neurons. Although the impact of NeuroD1 expression levels and duration on the functionality of the converted neurons remains to be determined, we observed that NeuroD1-converted cells did not develop the functional properties of mature neurons even after 17 weeks post-NeuroD1 induction. Thus, optimizing the expression timing, regional targeting, and cell-state specificity of neurogenic factors is critical for promoting functional maturity and integration of these converted neurons, ultimately enhancing their therapeutic efficacy for CNS injuries and neurodegenerative diseases. MATERIALS AND METHODS Animals NeuroD1, NeuroD1(1 to 155 amino acids), and Sox2-inducible knock-in mice were crossed with Aldh1l1-CreERT2 mice (JAX strain no. 031008) to generate iNeuroD1, iNeuroD1-mut, and iSox2 mice. Rosa-tdTomato mice (Ai9, JAX strain no. 007909) were crossed with Aldh1l1-CreERT2 mice to generate Control (Ai9; Aldhl1l-CreERT2) mice. Both conditional single and double transgenic mice were used in our experiments. Both male and female mice were used in this study, with no sex-specific phenotypic differences observed. All mice used in experiments were maintained in a pathogen-free vivarium with a 12-hour light/dark cycle with free access to normal chow food and water. All animal experiments were approved by the Institutional Animal Care and Use Committee of Children’s Hospital of Fudan University and Cincinnati Children’s Hospital Medical Center ([2025]62). Generation of inducible transgenic knock-in mice by CRISPR-Cas9 system The CAG-loxP-stop-loxP-NeuroD1 or NeuroD1(1 to 155 amino acids)–EGFP–WPRE (Woodchuck Hepatitis Virus Posttranscriptional Regulatory Element) were inserted into the Rosa locus. Biocytogen’s CRISPR-Cas9–based EGE system was used to prepare gene knock-in mice. Single guide RNAs (sgRNAs) are designed near the genomic insertion site, with targeting vectors’ 5′ and 3′ homology arms being 1812 base pairs (bp) and 1779 bp each. Guide RNAs were screened for on-target activity use UCA (Universal CRISPR Activity Assay), an sgRNA activity detection system developed by Biocytogen, and validated sgRNAs and Cas9 were transcribed in vitro. Both the Cas9 mRNA and the sgRNAs were microinjected, mixed, and co-injected into the cytoplasm of one-cell-stage fertilized eggs. After injection, surviving zygotes were transferred into oviducts of Institute of Cancer Research (ICR) albino pseudopregnant females. To generate inducible iSox2^mCherry knock-in mice, a donor vector was constructed in which the mCherry-Sox2-SV40 pA was in the reverse orientation and flanked by Lox66 and Lox71, allowing the one-way inversion upon expression of Cre. The transgene was put under the CAG promoter and protected by an insulator from the unwanted promoter activity on the 3′-side. The vector was targeted to the Col1a1 safe harbor locus with the aid of Cas9-gRNA that cleaves the target site. Pups were born and genotyped by PCR screening. Progeny from three founder mice carrying individual transgene gave rise to the same phenotype. The data presented are derived from the progeny of a single transgenic knock-in line. Spinal cord injury In our experiments, we performed the dorsal hemisection SCI model in adult mice (~8 to 10 weeks). All mice were anesthetized with 4% isoflurane until unconscious followed by 2% isoflurane during surgery. After exposing the spinal vertebrae at the level of T9-T12, meningeal tissue in the intervertebral space was cleared to minimize fibroblast invasion into the lesion core. Notably, we did not perform the laminectomy but only separated the intervertebral space for exposure to reduce post-injury tissue adhesion. Then, the hemisection was created from midline to lateral side by using a Sharpoint Stab Knife (15°, straight), extending toward the ventral surface of the cord. After the surgery, all mice were placed back to cages equipped with clean bedding and 37°C heating blanket to await recovery. The date of injury was denoted as dpi 0. Stab wound injury In our experiments, we performed the stab wound model in adult mice (~8 to 10 weeks). All mice were anesthetized with 4% isoflurane until unconscious followed by 2% isoflurane during surgery. An incision was made along the midline of the scalp, the periosteum was cleared, and bregma was exposed. A 1-mm hole was drilled into the skull through a 1-mm aperture centered 1 mm lateral from bregma. Stab wounds were placed into the somatosensory cortex (bregma from −0.9 to −2.7 mm, latero-lateral 1 mm) using Sharpoint Stab Knife. After the surgery, all mice were placed back to cages equipped with clean bedding and 37°C heating blanket to await recovery. The date of injury was denoted as dpi 0. Photothrombotic stroke Photothrombotic stroke injury was performed in adult mice (~8 to 10 weeks). All mice were anesthetized with 4% isoflurane until unconscious followed by 2% isoflurane during surgery. An incision was made along the midline of the scalp, the periosteum was cleared, and bregma was exposed. The skull was illuminated with 532-nm laser for 12 min through a 3-mm aperture centered 2 mm lateral from bregma beginning 5 min after intraperitoneal injection of rose bengal (15 mg/ml) into mice at 150 mg/g. After the surgery, all mice were placed back to cages equipped with clean bedding and 37°C heating blanket to await recovery. The date of injury was denoted as dpi 0. TAM and BrdU administration TAM (Sigma-Aldrich, T5648) was dissolved in 100% ethanol and then diluted in Core oil (MedChemExpress, HY-Y1888) to a final concentration of 20 mg/ml. TAM was intraperitoneally injected into mice at 100 mg/kg (~100 μl) once per day. BrdU (Sigma-Aldrich, B5002) was dissolved in double-distilled water in 10 mg/ml and preserved at −20°C from light. Consist BrdU labeling was performed by intraperitoneally injecting BrdU (50 mg/kg) for labeling reactive proliferative glia before the TAM injection. The time points of injections for each experiment are described in Results, illustration in figures, or figure legends. Tissue processing and immunohistochemistry Mice were anesthetized with Avertin and perfused with phosphate-buffered saline (PBS) followed by 4% paraformaldehyde (PFA). Brains and spinal cords were isolated and post-fixed for 2 hours with 4% PFA at 4°C and dehydrated in 25% sucrose at 4°C until they were completely settled. Samples were frozen in Tissue-Tek optimum cutting temperature compound (Sakura, catalog no. 4583) with dry ice and cryosectioned 10 μm thick using a sliding microtome (Leica). For immunostaining, cryosections were blocked with blocking solution (5% normal donkey serum and 0.2% Triton X-100) for 1 hour at room temperature (RT). The sections were then incubated with primary antibodies in the same blocking solution at 4°C overnight. For BrdU staining, cryosections were preprocessed with heat-induced antigen retrieval for 10 min at 95°C in sodium citrate buffer (10 mM, pH 6) and cooled for 15 min at RT. The primary antibodies used for immunofluorescence were goat anti-GFP (Novus Biologicals, catalog no. NB100-1770; RRID: AB_523903; 1:1000), chicken anti-GFAP (Millipore, catalog no. AB5541; RRID: AB_177521; 1:1000), rabbit anti-NeuN (Millipore, catalog no. ABN78; RRID: AB_10807945; 1:400), rabbit anti-NeuroD1 (Abcam, catalog no. ab109224; RRID: AB_10861489; 1:200), rabbit anti-OLIG2 (Abcam, catalog no. AB9610; RRID:AB_570666; 1:1000), goat anti-OLIG2 (NOVUS, catalog no. AF2418; RRID:AB_2157554; 1:40), rat anti-PDGFRα (BD Biosciences, catalog no. 558774; RRID: AB_397117; 1:100), rabbit anti-Iba1 (FUJIFILM Wako Shibayagi, catalog no. 019-19741; RRID: AB_839504; 1:1000), rat anti-BrdU (Abcam, catalog no. ab6326; RRID: AB_305426; 1:500), rabbit anti-S100β (Proteintech, catalog no. 15146-1-AP; 1:200), rabbit anti-DCX (Cell Signaling Technology, catalog no. 4604; RRID: AB_2276960; 1:800), rabbit anti-MAP2 (Abcam, catalog no. ab35454; RRID: AB_776174; 1:1000), mouse anti-vGlut2 (Abcam, catalog no. ab79157; RRID: AB_1603114; 1:500), and chicken anti-vGAT (Synaptic Systems, catalog no. 131006; RRID: AB_887872, 1:500). After brief washing with PBS (for 5 min, three times), the sections were incubated with secondary antibodies for 2 hours at RT. The secondary antibodies used in this study were Alexa Fluor Cy2-, Cy3-, or Cy5- conjugated donkey immunoglobulin G (IgG) against rat, rabbit, mouse, goat, chicken (Jackson ImmunoResearch; 1:500). Nuclei were co-stained with 4′,6-diamidino-2-phenylindole. Images were captured using confocal microscopy (Leica TSC SP8). Western blot analysis Tissues were isolated and homogenized in radioimmunoprecipitation assay buffer (Cell Signaling Technology, catalog no. 9806) supplemented with protease inhibitor cocktail (Roche, catalog no. 11873580001), phosphatase inhibitor (Roche, catalog no. 4906837001), and phenylmethylsulfonyl fluoride (Sangon Biotech, catalog no. 1861274). Lysates were quantified with the Beyotime’s BCA Protein Assay Kit. Equal amounts of these lysates (25 mg per lane) were resolved with a 12% bis-tris gel, transferred to a 0.45-μm polyvinylidene difluoride membrane (EMD Millipore, catalog no. HATF00010) by the electrophoretic transfer method. The membranes were blocked in 5% nonfat powdered milk [in 0.2% TBST (Tris-buffered saline with Tween 20)] for 1 hour and incubated with primary antibody at 4°C overnight. The primary antibodies used in this study were rabbit anti-NeuroD1 (Abcam, catalog no. ab109224; 1:1000) and mouse anti–β-actin (ORIGENE, catalog no. TA811000; 1:2000). After washing the blot membrane with 0.2%TBST (10 min, three times), the membranes were incubated with horseradish peroxidase–conjugated anti-rabbit or anti-mouse IgG (Jackson ImmunoResearch; 1:20,000) in 0.2%TBST for 1 hour at RT. After washing with 0.2%TBST (10 min, three times), the membranes were developed with chemiluminescence reaction using an ECL kit (Millipore, catalog no. WBKLS0500) and signal detection using ChemiDoc XRS+ imaging system (Bio-Rad). Acute brain slice processing The acute brain slices were prepared as previously described ([217]72). Mice were anesthetized with pentobarbital sodium (100 mg/kg, intraperitoneally) and perfused transcardially with ice-cold slicing solution, and the spinal cords were rapidly dissected and chilled in ice-cold sucrose–artificial cerebrospinal fluid (ACSF) solution. The tissues were placed on a previously trimmed 3% agar block and attached the ventral side with superglue to the agar in a vertical direction. Parasagittal slices (200 μm) were sliced with a speed of 0.12 mm/s using a vibratome (VT 1200 s, Leica). The slices were transferred to warm (34.5°C) ACSF composed of 126 mM NaCl, 2.5 mM KCl, 26 mM NaHCO[3], 1.25 mM NaH[2]PO[4], 2 mM MgSO[4], 2 mM CaCl[2], and 25 mM dextrose (290 mosmol, pH 7.2) and were bubbled with 95% O[2] and 5% CO[2]. For whole-cell patch-clamp recordings, slices were transferred to a submerged chamber on the infrared-differential interference contrast microscope (BX-51WI, Olympus). We used a multiclamp 700B amplifier (Molecular Devices) and a Power1401 for whole-cell recording and data acquisition. GFP^+ cells were identified using transmitted light and green fluorescence. During the recordings, borosilicate glass pipettes (4 to 7 megohms) were filled with an internal solution containing 130 mM K-gluconate, 10 mM KCl, 4 mM MgATP, 0.3 mM Na[2]GTP, 10 mM Hepes, 10 mM Na[2]-phosphocreatine, and 0.2% biocytin (288 mosmol, pH 7.29) or 140 mM K-gluconate, 3 mM KCl, 2 mM MgCl[2], 10 mM Hepes, 0.2 mM EGTA, 2 mM Na[2]ATP, and 0.2% biocytin (285 to 295 mosmol, pH 7.2). To examine the electrophysiological properties, positive current pulses (10 pA per step, 500 ms in duration) were injected to the cell recorded when it was at the resting membrane potential. To obtain the Na^+ currents, we obtained whole-cell recording with patch pipettes filled with Cs^+-based internal solution containing 140 mM CsCl, 2 mM MgCl[2], 2 mM Na[2]ATP, 10 mM Hepes, and 10 mM EGTA (287 mosmol, pH 7.2 with CsOH). CdCl[2] (100 μM) was included in the bath solution to block voltage-gated Ca^2+ currents. Cells were held at −100 mV for 80 ms and then depolarized to different membrane potential levels (from −90 to 60 mV, 10 mV per step) for 50 ms. P/N subtractions were performed offline by using Signal software. Spike2 version 10 and Signal software version 8 (Cambridge Electronic Design) together with MATLAB (MathWorks R2021a) were used for data analysis. In this study, spikelet was defined by three criteria: (i) the rising phase must reach a slope of at least 20 V/s (that defines the voltage threshold), (ii) the voltage difference between the peak and threshold must exceed 30 mV, and (iii) the absolute peak voltage must surpass 0 mV. The smaller spikelet, representing a more immature action potential variant, retained the same threshold slope and peak voltage requirements but differed in two aspects: the peak-to-threshold voltage difference was reduced to 10 to 30 mV, and threshold current (rheobase) exceeded 4500 pA when using 500-ms current injection. Single-nucleus isolation and sequencing Spinal cords were rapidly extruded from iNeuroD1^GFP mice (n = 3) after anesthesia. Each cord was separated into lesion core and peri-injury regions. snRNA-seq analysis was performed by the commercial service of Genergy Biotechnology Co. Ltd. (Shanghai, China). Nuclei were loaded onto the 10X Chromium Single Cell Platform (10X Genomics) at a concentration of 700 to 1200 nuclei per microliter (Single Cell 3′ library and Gel Bead Kit v.3) as described in the manufacturer’s protocol. Generation of gel beads in emulsion (GEMs), barcoding, GEM-RT clean-up, complementary DNA amplification, and library construction were all performed as per the manufacturer’s protocol. Qubit was used for library quantification before pooling. The final library pool was sequenced on the Illumina NovaSeq X plus instrument using 150-bp paired-end reads. Single-nucleus RNA-seq analysis Mapping to mouse mm10 genome, quality control, and read counting of Ensembl genes were performed by cellranger software with default parameter (v7.1.0). Unsupervised clustering and visualization were performed with R (Seurat package version 2.2). Genes expressed in fewer than two cells were filtered out. Cells with >200 genes and <10% mitochondrial genes were further processed. Then, variation coefficient of genes was calculated with Seurat. Dimensionality reduction of data was performed by using component analysis based on the first 2000 highest variable genes. A k-nearest neighbor graph was constructed from Euclidean distances in the space of the first 10 significant principal components. Louvain modularity optimization algorithm was used to cluster the cells in the graph, and clustering results were visualized by using UMAP. Cells expressing high levels of genes encoding hemoglobin were discarded. Differential expression of each cluster was calculated using the “bimod” test as implemented in Seurat FindMarkers function. Seurat-Bimod statistical test was used to find differentially expressed genes between each group of cells and other groups of cells [False Discovery Rate (FDR) ≤ 0.05 and |log2 fold change| ≥ 1.5]. GO enrichment analysis for these significant differentially expressed genes was performed by TopGO R package, and the Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway enrichment analysis was performed using the Hypergeometric test in R. Significantly enriched GO terms and KEGG pathways were selected by a threshold FDR (adjusted P value) ≤ 0.05. Cell clusters were annotated using canonical markers of known cell types. Quantification and statistical analysis All analyses were conducted using GraphPad Prism v10 (San Diego, California, [218]www.graphpad.com). Data are presented as means ± SEM. Statistical significance between two datasets was determined using a two-tailed Student’s t test. For comparisons involving multiple groups, one-way analysis of variance (ANOVA) followed by a post hoc multiple comparisons test was performed. Statistical significant thresholds were set at *P < 0.05, **P < 0.01, and ***P < 0.001. While no randomization method was applied for data collection, all data were quantified in a blinded manner. Acknowledgments