Abstract PALS1-associated tight junction (PATJ) protein is linked to metabolic disease and stroke in human genetic studies. Despite the recognized role of PATJ in cell polarization, its specific functions in metabolic disease and ischemic stroke recovery remain largely unexplored. We explored the functions of PATJ in an in vitro model and in vivo in C. elegans and mice. Using a mouse model of stroke, we found post-ischemic stroke duration-dependent increase of PATJ abundance in endothelial cells. PATJ knock-out (KO) HEK293 cells generated by CRISPR-Cas9 suggest roles for PATJ in cell proliferation, migration, mitochondrial stress response, and interactions with the Yes-associated protein (YAP)-1 signaling pathway. Notably, PATJ deletion altered YAP1 nuclear translocation. PATJ KO cells demonstrated transcriptional reprogramming based on RNA sequencing analysis, and identified dysregulation in genes central to vascular development, stress response, and metabolism, including RUNX1, HEY1, NUPR1, and HK2. Furthermore, we found that mpz-1, the homolog of PATJ, was significantly upregulated under hypoxic conditions in C. elegans. Knockdown of mpz-1 resulted in abnormal neuronal morphology and increased mortality, both of which were exacerbated by hypoxia exposure, indicating a critical protective role of PATJ/MPZ-1 in maintaining neuronal integrity and survival, particularly during oxygen deprivation stress relevant to ischemic stroke. These insights offer a new understanding of PATJ's regulatory functions within cellular and vascular physiology and help lay the groundwork for therapeutic strategies targeting PATJ-mediated pathways for stroke rehabilitation and neurovascular repair. Keywords: PATJ, Ischemic stroke, Vascular remodeling, Cellular stress, YAP1 Graphical abstract [33]Image 1 [34]Open in a new tab Highlights * • PATJ modulates hypoxic stress responses in C. elegans and mammalian cells, suggesting a role in ischemic injury resilience. * • PATJ expression in brain endothelial cells is dynamically regulated following ischemic stroke, suggesting its involvement in vascular repair processes. * • PATJ deletion induces nuclear translocation of YAP1, indicating a novel regulatory interplay with the Hippo pathway. * • Transcriptomic analysis reveals PATJ-dependent regulation of genes and pathways critical for vascular development, stress responses, and metabolism. 1. Introduction Stroke is a principal cause of mortality and prolonged disability in the United States. While thrombolysis with tissue-type plasminogen activator (tPA) or endovascular thrombectomy has improved outcomes following acute ischemic stroke (AIS) [[35]1], strategies to improve outcomes for stroke survivors are still notably lacking [[36]2]. Angiogenesis, the process of new blood vessel formation from existing ones is tightly regulated under normal physiological conditions in adults [[37]3]. Following AIS, angiogenesis is a critical adaptation restoring cerebral blood flow and supporting repair processes [[38][4], [39][5], [40][6]]. Interestingly, an increased density of microvessels in brain regions impacted by AIS correlates with better prognosis in patients [[41]7], thus promoting angiogenesis following AIS is a promising method for enhancing neurorecovery from AIS. The evolutionarily conserved Crumbs protein homologs (CRB) are key components in scaffolding and cellular orientation. Central to this orchestration is the PALS1-associated tight junction (PATJ) protein, a component of the CRB complex. PATJ serves as a connector between the CRB–PALS1 complex and tight junction proteins [[42]8,[43]9], maintaining cell polarization and ensuring the stability of apical junctions [[44]10,[45]11]. However, recent advancements suggest a broader spectrum of PATJ's functionality including roles in signal transduction, cellular dynamics, and tumorigenesis [[46]12]. A recent genome-wide association study (GWAS) on acute ischemic stroke (AIS) outcomes revealed that AG or GG single nucleotide polymorphisms (SNPs) of the PATJ gene are strongly associated with poor 90-day modified Rankin Scale (mRS) outcomes post-AIS [[47]13]. Interestingly, SNPs in PATJ have also been implicated in cardiometabolic health and sleep disturbances [[48]14,[49]15]. These genome-scale findings provide genomic links between PATJ and cardiovascular pathology, suggesting that variations in PATJ not only impact cellular polarity but may also influence cardiovascular risk and AIS recovery outcomes. PATJ knockdown in human microvascular endothelial cells was found to promote endothelial to mesenchymal transition and pro-angiogenic programs [[50]16]. Prior work indicates an important role for PATJ in directing embryonic capillary migration in cooperation with angiomotin (AMOT) and Syx, a RhoA GTPase exchange factor (RhoGEF) protein also known as Pleckstrin homology domain containing, family G member (PLEKHG)-5 [[51]17]. The involvement of endothelial cells in vascular remodeling and angiogenesis is well-documented, yet the specific contributions of PATJ during post-AIS angiogenesis remains incompletely understood. Our study builds upon this recent expanded understanding of PATJ, and links PATJ with post-AIS angiogenesis, cellular migration, proliferation, and metabolism. We also found regulatory interplay between PATJ and key signaling pathways, notably YAP1 within the Hippo pathway. Taken together, these findings highlight PATJ's role in neurorepair-related pathways and suggest its potential as a novel therapeutic target. 2. Methods 2.1. CRISPR-Cas9 RNP complex preparation and cell transfection For the targeted editing of PATJ exon 3, single guilde (sg)RNA sequences sgRNA1 (CUUGACCCAAGAUAAACUGC) and sgRNA2 (GUGAGUAUCUGGUUGAAGAG) were synthesized with high specificity and minimal off-target effects (Synthego). The sgRNAs were then complexed with Cas9 protein (IDT) to form ribonucleoprotein (RNP) complexes. The complexes were prepared by diluting sgRNA and Cas9 to 3 μM in Opti-MEM™ I Reduced Serum Medium, adhering to a sgRNA:Cas9 molar ratio of 1.3:1, followed by a brief incubation. Human embryonic kidney (HEK)293 cells, cultured in Dulbecco's Modified Eagle Medium (DMEM) supplemented with 10 % Fetal Bovine Serum and antibiotics, were grown to 30–70 % confluence in 24-well plates to ensure optimal cell density for transfection. The transfection utilized Lipofectamine™ CRISPRMAX™ (Invitrogen) for RNP complex delivery, employing a reverse transfection technique where the RNP solution is pre-mixed with the cell suspension before plating. Following transfection, the cells were incubated at 37 °C and 5 % CO2, with medium changes to maintain cell health and promote recovery. 2.2. Isolation and validation of CRISPR-edited clones Following transfection, single cell derived clones were derived from singly plated cells, ensuring genetic uniformity. Over 2–8 weeks, these colonies were observed under standard conditions and expanded based on visual confirmation via microscopy. Successfully expanded colonies were then transferred to 24-well plates for further growth. Western blot was performed to validate protein knockdown. Sanger sequencing analysis post-PCR amplification from genomic DNA of individual clones revealed insertion/deletion in the region of PATJ exon 3 targeted by sgRNA1. 2.3. Migration assay Confluent HEK293 cell monolayers, including wild-type (WT), PATJ knockout (KO), PATJ overexpression (Ov-PATJ), and KO with PATJ overexpression (KO + Ov-PATJ), were subjected to a scratch assay. PATJ overexpression was achieved by transfecting cells with a plasmid encoding full-length human PATJ using Lipofectamine 2000 (Invitrogen) according to the manufacturer's instructions. An empty vector was used as a control. Where indicated, cells were treated with verteporfin (1 μM) to inhibit YAP1 activity. A sterile pipette tip was used to create a straight 'wound' across the monolayer. Cell proliferation was inhibited using Mitomycin C (10 μg/mL). The migration of cells into the wound area was imaged at specified intervals to assess closure rate. Quantitative analysis was conducted by measuring the gap width at 0-, 24-, and 48-h post-scratch. Images were captured under a Zeiss AXIO Observer.A1 Inverted Fluorescence Microscope. Cellular migration was determined by ImageJ software. Relative migration rate was calculated as (Areawound at 0 h – Areawound at 24 h)/Areawound at 0 h. 2.4. Rotenone-induced cell growth assessment in PATJ-modified HEK293 cells HEK293 cells, both with and without PATJ knockout, were cultured in 24-well plates at a seeding density of 10^5 cells per well. For cell proliferation and survival assays PATJ KO and WT cells were cultured under normal conditions (Vehicle), with rotenone treatment (200 nM), or with combined rotenone (200 nM) and verteporfin (1 μM) treatment. Verteporfin is a specific inhibitor of YAP1 activity. The rotenone concentration was selected to inhibit mitochondrial complex I, thereby simulating metabolic stress analogous to ischemic conditions, while verteporfin was used to determine whether YAP1 inhibition could modulate the observed effects of PATJ knockout on rotenone sensitivity. Cell proliferation was assayed by automated cytometry quantification of cell number at specified intervals post-treatment (Nanoentek, Waltham, MA, USA). Propidium iodide staining followed by automated cytometry was performed for assessment of cell death following rotenone (2 μM). 2.5. Seahorse XF metabolic flux analysis Mitochondrial respiration in HEK293 cells, with and without PATJ modification, was evaluated using the Seahorse XF^e96 Extracellular Flux Analyzer (Seahorse Bioscience, North Billerica, MA, USA) with the Seahorse XF Cell Mito Stress Test Kit. Cells were plated at 25,000 cells per well in XF^e96 microplates and incubated in standard culture medium. Following overnight incubation, cells were washed and equilibrated in unbuffered Seahorse XF Base Medium supplemented with glucose (4.5 g/L), l-glutamine (2 mM), and sodium pyruvate (1 mM) in a non-CO2 incubator at 37 °C for 1 h. Oxygen consumption rate (OCR) was recorded before and after sequential injections of mitochondrial inhibitors: oligomycin (1 μM), Trifluoromethoxy carbonylcyanide phenylhydrazone (FCCP; 1 μM), and rotenone/antimycin A (1 μM). Post-assay, the cells stained with Hoechst 33258 (0.1 μg/mL) and fluorescence intensity acquired at excitation/emission 355/460 nm to normalize the OCR data. 2.6. Assessment of stress responses in C. elegans C. elegans strains were maintained on nematode growth medium (NGM) plates seeded with Escherichia coli OP50 at 20 °C using standard procedures [[52]18]. The mpz-1(mib188[mpz-1::gfp]) II strain [[53]19] was used in this study. For stress treatments, L4 stage worms were subjected to either hypoxia (0.1 % O2) or normoxia for 72 h, or heat stress at 20 °C (control), 25 °C, or 28 °C for 72 h [[54]20]. For confocal imaging, treated MPZ-1GFP worms were randomly selected, immobilized with 10 mM sodium azide in M9 buffer, and aligned on 2 % agarose pads on glass slides. Images were acquired using a Leica TCS SPE confocal microscope with a 63 × objective lens, maintaining consistent settings across all samples. MPZ-1GFP abundance was quantified using ImageJ. 2.7. C. elegans strains and RNAi treatment C. elegans strains were maintained on nematode growth medium (NGM) plates seeded with Escherichia coli OP50 at 20 °C using standard procedures. For high-resolution single-neuron morphological analysis, we utilized the otIs181[Pdat-1mCherry; Pttx-3mCherry] strain [[55]21] crossed with the uIs69 [(pCFJ90) myo-2pmCherry + unc-119psid-1] strain [[56]22]. This combined strain specifically labels dopaminergic cephalic neurons (CEP) and AIY interneurons in the head region while providing enhanced neuronal sensitivity to RNAi treatment, enabling effective assessment of gene knockdown effects specifically within the labeled neuronal populations. RNAi knockdown of mpz-1 was performed by feeding worms with HT115 bacteria expressing double-stranded RNA targeting mpz-1 from the embryonic stage. Empty vector (L4440) was used as a control. Day 3 adult worms were subjected to 72 h of hypoxia (0.1 % O2) or normoxia. Worm survival was assessed by touch-provoked movement. Worms unresponsive to gentle prodding with a platinum wire were scored as dead. For confocal imaging, treated worms were randomly selected, immobilized with 10 mM sodium azide in M9 buffer, and aligned on 2 % agarose pads on glass slides. Images were acquired using a Leica TCS SPE confocal microscope with a 63 × objective lens, maintaining consistent settings across all samples. Individual neurons were classified as "abnormal" if they exhibited aberrant branching patterns, process fragmentation, blebbing, or irregular axonal/dendritic morphology compared to control neurons. The percentage of abnormal worms with abnormal mCherry-labeled neurons was calculated as the ratio of worms with neuronal abnormalities to the total number of worms examined in each condition. 2.8. Western blot Protein lysates were extracted from HEK293 cells, both PATJ KO and WT, using cell lysis buffer (Cell Signaling Technology, Danvers, MA) supplemented with Protease Inhibitor Cocktail (Cell Signaling Technology, Danvers, MA). Equal amounts of protein were separated on 4–15 % precast gels (Bio-Rad) and then transferred onto polyvinylidene difluoride (PVDF, Bio-Rad) membranes. Membranes were blocked for 1 h at room temperature using Tris-buffered saline with 0.1 % Tween-20 (TBS-T) supplemented with 5 % non-fat dried milk. Membranes were then incubated overnight at 4 °C with primary antibodies diluted in TBS-T containing 5 % bovine serum albumin (BSA). Primary antibodies used were PATJ (1:1000, Proteintech) and YAP1(1:2000, Proteintech). Subsequently, membranes were probed with appropriate HRP-conjugated secondary antibodies for 1 h at room temperature. Protein bands were visualized using an enhanced chemiluminescence (ECL) detection system (Azure Biosystems). Densitometric analysis was conducted to quantify protein expression levels, with α-tubulin employed as a loading control. 2.9. Cellular immunofluorescence HEK293 cells, both PATJ KO and WT, were cultured on coverslips until optimal confluence was reached. Cells were fixed in 4 % paraformaldehyde, then permeabilized with 0.1 % Triton X-100 in PBS for 15 min. Non-specific binding was minimized by blocking with goat serum. Overnight incubation at 4 °C with primary antibodies targeting YAP1 (1:200, Proteintech) was followed by application of goat anti-rabbit Alexa Fluor 488 secondary antibody (1:1000, Invitrogen). For conventional fluorescence microscopy, nuclei were stained with DAPI. For confocal microscopy, the cytoskeleton was additionally stained with CoraLite 594-Phalloidin (red, Proteintech). Coverslips were mounted and examined using either a Zeiss AXIO Observer.A1 Inverted Fluorescence Microscope or a Leica confocal microscope. Coverslips were mounted for examination by Leica confocal microscopy, focusing on YAP localization and expression. YAP1 subcellular localization was quantified using two complementary approaches: (1) Distribution pattern analysis: YAP1 subcellular distribution was visually assessed in at least 50 cells per condition per experiment across four independent experiments using standard fluorescence microscopy. Cells were classified into three categories based on YAP1 distribution: predominantly nuclear, predominantly cytoplasmic, or distributed in both compartments. The percentages of cells in each category were calculated for each experimental condition. (2) Nuclear-to-cytoplasmic intensity ratio measurement: For quantitative assessment of YAP1 subcellular distribution, confocal microscopy images were analyzed using ImageJ software. Regions of interest (ROIs) were defined within the nuclear area (identified by DAPI staining) and in the surrounding cytoplasmic area of each cell. Mean fluorescence intensity of YAP1 was measured in these ROIs, and the nuclear-to-cytoplasmic ratio was calculated. A minimum of 50 cells per condition per experiment were analyzed across four independent experiments. All image acquisition parameters were kept constant between samples to ensure comparable measurements. 2.10. Mouse model of transient focal cerebral ischemia Transient focal cerebral ischemia was induced in C57BL6 mice aged 8–10 weeks (weighing 23–26 g) using the Middle Cerebral Artery Occlusion (MCAO) technique. Mice were anesthetized with isoflurane, delivered in a mixture of oxygen and nitrous oxide. A midline neck incision was made to expose the common carotid artery, followed by electrocoagulation of its branches. A nylon filament was carefully inserted through the external carotid artery to the internal carotid artery, and advanced towards the middle cerebral artery (MCA) to achieve a reduction in regional cerebral blood flow (CBF) to less than 25 % of the baseline values. Monitoring of the regional cerebral blood flow was conducted through Laser Doppler Flowmetry (Moor Instruments Ltd, UK). The mice's body temperature was maintained at 37.0 ± 0.5 °C using a regulated heating pad, monitored with a rectal thermometer throughout the procedures. The occlusion was sustained for 60 min, after which reperfusion was initiated. Mice were then allowed a recovery period ranging from 1 to 28 days. For the control group, sham operations were performed where all steps were identical except for the artery occlusion. Postoperative care included the administration of buprenorphine for pain relief. Mice failing to meet the specified CBF criteria or those that did not survive the procedure (n = 5) were excluded from subsequent analyses. 2.11. Immunohistochemistry After transient Middle Cerebral Artery Occlusion (tMCAO), mice were anesthetized deeply and perfused transcardially, first with 0.9 % NaCl, followed by 4 % paraformaldehyde in PBS. Brains were then extracted, cryoprotected in 30 % sucrose in PBS for 48 h, and sectioned coronally on a cryostat. Sections were preserved at −20 °C in a cryoprotectant solution until required for analysis or immunofluorescence staining. For staining, sections were washed in PBS, permeabilized with PBST (1 % Triton X-100 in PBS) for 20 min and blocked with 5 % normal goat serum for an hour. Following further washing with 0.3 % PBST, they were incubated overnight at 4 °C with rat anti-mouse CD31 (1:200, BD Pharmingen) and rabbit polyclonal PATJ antibody (1:100, LSBio). Alexa Fluor 488 conjugated goat anti-rat (1:1000, Invitrogen) or Alexa Fluor 594 goat anti-rabbit (1:1000, Invitrogen) secondary antibodies were applied respectively. For NeuN/PATJ staining, sections underwent incubation with NeuN (1:1000, LSBio) and PATJ antibodies, followed by the appropriate Alexa Fluor-conjugated secondary antibodies. Immunofluorescence images were captured using a Zeiss AXIO Observer.A1 Inverted Fluorescence Microscope. For YAP1/CD31/DAPI triple staining, sections were incubated overnight at 4 °C with rat anti-mouse CD31 (1:200, BD Pharmingen) and rabbit polyclonal YAP1 antibody (1:200, Proteintech). Alexa Fluor 594 conjugated goat anti-rat (1:1000, Invitrogen) and Alexa Fluor 488 goat anti-rabbit (1:1000, Invitrogen) secondary antibodies were applied respectively. Sections were then mounted with VECTASHIELD Vibrance® Antifade Mounting Medium with DAPI (H-1800) for nuclear counterstaining. For triple immunofluorescence staining of CD31/YAP1/PATJ, sections were incubated overnight at 4 °C with rat anti-mouse CD31 (1:200, BD Pharmingen), mouse monoclonal YAP1 antibody (1:200, Proteintech), and rabbit polyclonal PATJ antibody (1:100, LSBio). Prior to applying the mouse primary antibody, sections were treated with a mouse-on-mouse blocking kit (Vector Laboratories) according to the manufacturer's instructions to minimize non-specific binding. After washing, sections were incubated with Alexa Fluor 594 conjugated goat anti-rat for CD31, Alexa Fluor 488 goat anti-mouse for YAP1, and Alexa Fluor 405 goat anti-rabbit for PATJ secondary antibodies (all 1:1000, Invitrogen) for 1 h at room temperature. Images were acquired using a CSU-W1 SORA spinning disk confocal microscope. 2.12. RNA-seq analysis of HEK293 cell lines Total RNA was extracted from HEK293 cells across three experimental groups: PATJ KO, WT, and PATJ KO with overexpression, utilizing the RNeasy Mini Kit (Qiagen). Prior to library preparation, the integrity and concentration of RNA were verified. Sequencing libraries were constructed following cDNA synthesis, adaptor ligation, and amplification. Library quality control followed by next-generation sequencing was performed on the DNBSeq platform by a commercial provider (BGI Americas, San Jose, CA). Reads were aligned to the human genome (Hgv38) for differential gene expression analysis using the limma R package (version 3.40.6; [[57]23]), which utilizes generalized linear models. Volcano plots and heatmaps were generated with the ggplot2 (version 3.4.0; [[58]24]) and pheatmap (version 1.0.12) R packages, respectively. For dimensionality reduction, the umap R package (version 0.2.7.0) was used, normalizing expression profiles (z-score) for UMAP analysis to generate a reduced-dimensionality matrix. Gene set enrichment analysis for both KEGG pathways and GO annotations was performed using the latest annotations from the KEGG REST API and org.Hs.eg.db R package (version 3.15.0), respectively. The clusterProfiler R package (version 4.4.4) mapped genes to these annotations, identifying statistically significant pathways and GO terms. Enrichment analyses considered a minimum gene set of 5, a maximum of 5000, and an FDR threshold of <0.05 as significant. 2.13. Bioinformatics analysis of PATJ expression in hypoxic endothelial cells Bioinformatics analyses using transcriptomic data sets available on Gene Expression Omnibus (GEO) to complement the present RNA sequencing. Six GEO RNA-seq datasets ([59]GSE107029, [60]GSE76743, [61]GSE70330, [62]GSE145774, [63]GSE186616, [64]GSE71216) were identified in human umbilical vein endothelial cells (HUVEC) exposed to hypoxic and/or control conditions. The samples consisted of 14 hypoxia (0.2 %–1 %, 24-h treatment) and 14 normoxia treated samples. Raw FPKM data were downloaded, and batch effects were removed using the ComBat method from the "sva" R package. Principal component analysis (PCA) was used to evaluate batch effect removal. PATJ expression levels were compared between normoxic and hypoxic conditions using boxplots, with statistical significance assessed by Wilcoxon rank-sum test. For single-gene GSEA analysis, Pearson correlation between PATJ and all other genes was calculated using the "cor" function in R, and genes were ranked by correlation coefficient. The "GSEA" function from the "clusterProfiler" package was used with KEGG pathways as predefined gene sets. Pathways with |normalized enrichment score (NES)| >1 and p < 0.05 were considered significant. Correlation analyses between PATJ and genes involved in Hippo signaling, angiogenesis, and cell proliferation/migration were conducted using Pearson's correlation method. Correlation heatmaps and scatterplots were generated to visualize relationships between PATJ and functional genes of interest. 2.14. Statistical analysis All statistical analyses were performed using GraphPad Prism version 10 or R. Data distributions were first assessed for normality using the Shapiro-Wilk test. Group comparisons were conducted using one-way or two-way analysis of variance (ANOVA), followed by Bonferroni post hoc correction for multiple comparisons. Comparisons between two groups were conducted using two-tailed unpaired Student's t-tests for normally distributed data with equal variances. Statistical significance was defined as p < 0.05. All data are presented as mean ± standard error of the mean (SEM). 3. Results 3.1. Human and mouse brain transcriptomic atlases reveal developmental trajectory and cell subtype specific expression patterns of PATJ To extend prior human genetic work implicating PATJ in metabolic disease and stroke and provide further insights into PATJ function, we examined existing large mouse and human brain single cell RNA sequencing atlases profiling PATJ expression across neural cell types and the lifespan [[65][25], [66][26], [67][27]]. Interestingly, during fetal development and infancy PATJ is particularly enriched and levels decrease into adulthood. This pattern of expression is seen in many other key genes involved in neurogenesis and vasculogenesis such as ephrin B1, paired box protein (PAX)-6, neuronal differentiation (NeuroD)-1, and beta-catenin (CTNNB1), among others ([68]Fig. 1A). Although PATJ is enriched in early life, its expression in a broad array of cells is still present into adulthood in human and mouse single cell RNA atlas data ([69]Fig. 1B–C). In particular, single cell RNA sequencing atlas of whole human or mouse brain demonstrates that PATJ expression is most prominent in cerebellar glutamatergic cells, but also highly expressed in endothelial and other brain vascular cells (particularly those of the choroid plexus; [70]Fig. 1B–C). Of note, the cell-specific distribution of other key genes previously suggested to interact with PATJ including AMOT, PLEKHG5, and YAP1 demonstrate similar enrichment in cerebral vascular cell subtypes (see [71]Supplementary Fig. 1A–C). Fig. 1. [72]Fig. 1 [73]Open in a new tab PATJ expression across development, cell types and endothelial hypoxia. (A) Dotplot demonstrating expression of PATJ and other key regulators of neurogenesis and vasculogenesis in the cortex. PATJ expression is highest during fetal development and infancy and decreases in adolescence and adulthood. Data from Zhu et al., 2023 ([74]GSE204684) and accessed at [75]singlecell.broadinstitute.org. (B) Human and (C) Mouse single cell RNA sequencing brain atlas UMAP data demonstrating ubiquitous expression of PATJ in adult brain cell subtypes, with inset highlighting vascular cell subtypes. Data from Siletti et al., 2023 and Yao et al., 2023, respectively and accessed at [76]knowledge.brain-map.org. (D) Boxplot showing significantly increased PATJ expression in hypoxic versus normoxic HUVECs. ∗∗∗∗ p < 0.0001. (E) Gene-set enrichment analysis showing top pathways significantly associated with PATJ expression in HUVECs. We also analyzed PATJ RNA expression in existing GEO data sets of human umbilical vein endothelial cells (HUVECs) exposed to hypoxia (0.2–1.0 % for 24 h). After batch effect correction ([77]Supplementary Fig. 1D), we observed significant upregulation of PATJ expression in endothelial cells exposed to 24-h hypoxia compared to normoxia (p = 7.4 x 10^−7; [78]Fig. 1D). Co-expression gene set enrichment analysis revealed that PATJ-correlated genes in endothelial cells were significantly enriched in multiple pathways relevant to vascular remodeling, including Hippo signaling, PI3K-Akt signaling, MAPK signaling, extracellular matrix receptor interactions, and HIF-1 signaling ([79]Fig. 1E). Single-gene correlation analysis demonstrated strong positive associations between PATJ and key angiogenic factors, including VEGFA (r = 0.97), VEGFC (r = 0.94), and PECAM1 (r = 0.93) ([80]Supplementary Fig. 1E). Taken together, existing human and mouse cell type-specific gene expression data suggest functional roles for PATJ in both neuronal and vascular biology. Endothelial cell-specific analyses provide rationale for investigating PATJ's interactions with Hippo signaling and YAP1 and involvement in post-stroke vascular remodeling. 3.2. PATJ modulates cell proliferation and stress responses in mammalian cells and C. elegans CRISPR-Cas9 and a sgRNA targeting PATJ was used to achieve knock-out of PATJ in two clonal HEK293 cell lines ([81]Fig. 2A). Edited and non-edited cells were screened by Western blot and sequencing identified two clonal lines with frameshift mutations that resulted in early stop codons and absence of PATJ mRNA and protein ([82]Fig. 2B–C) as well as wild-type cells exposed to editing procedures but without successful editing. We found a statistically significant discrepancy in growth between PATJ KO and WT cells ([83]Fig. 2D). To investigate differences in metabolic responses related to PATJ deficiency, we used a Seahorse mitochondrial stress assay. Interestingly, differences in oxygen consumption were observed in vitro following administration of oligomycin and rotenone/antimycin ([84]Fig. 2E) while preserving spare respiratory capacity. The sensitivity to oligomycin, an inhibitor of ATP synthesis, and rotenone/antimycin, inhibitors of complex I and complex III, suggests a greater dependence on oxidative phosphorylation in PATJ KO cells. Rotenone, a mitochondrial complex I inhibitor, was used to model energetic stress in PATJ-deficient contexts. Rotenone treatment increased cell death in PATJ KO cells after 48 h compared to wildtype HEK293 cells ([85]Fig. 2F). This data collectively indicates that PATJ contributes to cellular fitness and adaptability in conditions of metabolic stress. Fig. 2. [86]Fig. 2 [87]Open in a new tab Generation and metabolic profiling of PATJ knock-out (KO) HEK293 cells. (A) Schematic representation of CRISPR-Cas9 strategy targeting exon 3 of the PATJ gene.) CRISPR-Cas9 introduced an insertion (KO1) or deletion (KO2) leading to a frameshift and premature stop codon (red underline). Lowercase sequence base pairs denote intronic sequence. Ex: exon. (B) RNA-seq gene counts demonstrate the absence of PATJ transcripts in KO versus WT cells, (n = 3 for each group). (C) Western blot verifying the absence of PATJ protein expression in knockout cell lines. (D) Growth comparison of PATJ knockout (KO) and wild-type (WT) HEK293 cells (n = 3 independent experiments with 3 technical replicates for each group). (E) Oxygen consumption rates of PATJ KO and WT cells following treatment with oligomycin (i), FCCP (ii), and antimycin A/rotenone (iii; n = 4–6 replicates for each group). (F) PATJ-KO cells demonstrate higher cell death as measured by propidium iodide staining than WT after 2 μM rotenone for 24 h (n = 3 independent experiments with 3 technical replicates for each group). ∗ p < 0.05, ∗∗ p < 0.01, ∗∗∗ p < 0.001, ∗∗∗∗ p < 0.0001. Prior work has implicated PATJ in developmental processes and suggested interaction with cytoskeletal proteins [[88]28]. As such we hypothesized that cell migration may be affected in PATJ KO cells as assessed by an in vitro scratch assay. In this assay a mechanical scratch is introduced into cells grown at high confluence and the resulting ‘scar’ monitored for cell migration and closure. Cell proliferation is inhibited with the addition of mitomycin C. We found a notable reduction in cellular migration in PATJ KO cells compared to WT HEK293, while accelerated migration was seen in cells overexpressing PATJ ([89]Fig. 3). Quantitative analysis demonstrated a significant increase in the gap width of PATJ KO cells at 24 h, which persisted to 48 h post-scratch. The migration impairment observed in PATJ KO cells was effectively rescued by overexpression of PATJ to levels comparable to WT cells. These results indicate a critical role for PATJ in cell migration following an injury. Fig. 3. [90]Fig. 3 [91]Open in a new tab Comparative analysis of cell migration in PATJ-modified HEK293 Cells. (A) Scratch assay time-lapse images at 0, 24, and 48 h post-wounding display cell migration for various HEK293 cell lines: wild-type (WT), PATJ knock-out (KO), PATJ overexpression (Ov-PATJ), and KO with PATJ overexpression (KO + Ov-PATJ). Dotted lines indicate scratch edges. (B) Quantitative analysis of cell migration at 24 h post-scratch. (C) Quantitative analysis of cell migration at 48 h post-scratch. n = 4 independent experiments with 3 technical replicates. ∗p < 0.05, ∗∗p < 0.01, ∗∗∗p < 0.001, ∗∗∗∗p < 0.0001. To investigate the evolutionary conservation of PATJ's role in stress responses, we examined the expression of MPZ-1, a C. elegans ortholog of PATJ, under hypoxic and heat stress conditions using a CRISPR-generated MPZ-1GFP reporter strain. This strain exhibits strong MPZ-1GFP expression in the head region of C. elegans ([92]Fig. 4A). Exposure to hypoxia (0.1 % O2) for 72 h resulted in a significant upregulation of MPZ-1GFP expression compared to normoxic controls ([93]Fig. 4B–C). In contrast, heat stress at 25 °C or 28 °C for 72 h did not significantly alter MPZ-1GFP levels relative to the 20 °C control ([94]Fig. 4D–E). These findings suggest that the induction of PATJ/MPZ-1 expression in response to hypoxia is specific and conserved in C. elegans. Fig. 4. [95]Fig. 4 [96]Open in a new tab PATJ modulates stress responses in C. elegans (A) Schematic diagram of C. elegans anatomy showing MPZ-1GFP is expressed in head neurons. (B) Representative confocal images of endogenous MPZ-1GFP expression in C. elegans under hypoxia (0.1 % O2) or normoxia for 72 h post L4 stages. Scale bars: 10 μm. (C) Quantification of MPZ-1GFP fluorescence intensity under hypoxic and normoxic conditions. n > 20 animals per condition. (D) Representative confocal images of MPZ-1GFP expression in C. elegans under heat stress at 20 °C, 25 °C, or 28 °C for 72 h post L4 stages. Scale bars: 10 μm. (E) Quantification of the percentage of worms showing enhanced MPZ-1GFP fluorescence intensity under different temperature conditions. Data points represent three independent plates with >20 animals assessed per condition. (F) Representative confocal images of individual neurons in otIs181[Pdat-1mCherry; Pttx-3mCherry] worms following control RNAi or mpz-1 RNAi treatment under normoxic and hypoxic conditions. Scale bars: 2 μm. (G) Quantification of the percentage of abnormal worms with abnormal neurons under different conditions. Data points represent three independent experiments with >20 animals assessed per condition. (H) Death rates of L4440 control and mpz-1 RNAi worms subjected to hypoxia for 72 h ∗p < 0.05, ∗∗p < 0.01, ∗∗∗∗p < 0.0001. To investigate the role of PATJ/MPZ-1 in neuronal health with single-cell resolution, we examined individual neuronal morphology using the otIs181[Pdat-1mCherry; Pttx-3mCherry] strain crossed with uIs69 strain, which specifically labels dopaminergic CEP and AIY interneurons in the head region ([97]Fig. 4F–G). This strain combination provides enhanced RNAi sensitivity while enabling precise visualization of individual neuronal morphology within defined neuronal populations. Under normoxic conditions, control RNAi-treated worms displayed normal neuronal morphology characterized by typical branching patterns and intact cellular integrity. In contrast, mpz-1 RNAi treatment induced noticeable morphological abnormalities in individual neurons, including irregular branching patterns and aberrant neuronal architecture. When subjected to hypoxic stress, these morphological defects were markedly exacerbated in mpz-1 RNAi-treated worms, with neurons displaying severe structural abnormalities, loss of normal branching patterns, and compromised cellular architecture. Although control animals also showed some sensitivity to hypoxic stress, the morphological changes were more pronounced following mpz-1 RNAi treatment. Notably, hypoxia exposure also led to a higher mortality rate in mpz-1 RNAi worms compared to control worms ([98]Fig. 4H). These results indicate that loss of PATJ/MPZ-1 function leads to neuronal defects and increased vulnerability to hypoxia-induced death, underscoring its protective role in maintaining neuronal integrity and survival, especially during oxygen deprivation stress relevant to ischemic stroke. 3.3. Interplay between PATJ and YAP1 in cellular stress and subcellular dynamics Given prior studies and bioinformatic evidence linking PATJ and YAP1 to the regulation of cellular architecture and stress responses through overlapping polarity and signaling pathways, we investigated PATJ and YAP1 functional interaction in mammalian cells. Western blot analysis demonstrated an upregulation in YAP1 protein levels in PATJ KO cells compared to WT ([99]Fig. 5A), suggesting a regulatory interplay between PATJ and YAP1 expression or stability. Immunofluorescence imaging revealed distinct differences in YAP1 subcellular distribution patterns between WT and PATJ KO cells ([100]Fig. 5B). Quantitative analysis showed that PATJ KO cells exhibited a significantly higher proportion of cells with predominantly nuclear YAP1 localization and a lower proportion with predominantly cytoplasmic localization compared to WT cells. Fig. 5. [101]Fig. 5 [102]Open in a new tab YAP1 expression and localization in PATJ-modified HEK293 cells and ischemic stroke.(A) Western blot analysis of YAP1 protein levels in WT and PATJ KO cells, with quantification showing a significant increase in YAP1 protein levels in KO cells (n = 4 for each group). (B) Immunofluorescence staining of YAP1 (green) DAPI staining of nuclei (blue). Yellow arrows indicate cells with predominantly cytoplasmic YAP1 localization, red arrowheads indicate predominantly nuclear YAP1 localization, and white arrowheads indicate cells with YAP1 distributed in both compartments. Right panels show quantitative analysis of YAP1 subcellular distribution patterns, presented as the percentage of cells with YAP1 localization predominantly in the nucleus, cytoplasm, or both compartments (n = 4 independent experiments with >50 cells/group). (C) Confocal microscopy with 3D reconstruction of YAP1 subcellular distribution in WT and PATJ KO cells. Right panel shows quantification of the YAP1 nuclear-to-cytoplasmic intensity ratio (n = 4 independent experiments with 50 cells/group). (D) Scratch-wound assay of WT and KO monolayers treated with verteporfin, captured at 0, 24, and 48 h post-scratch. Dashed lines denote wound edge (n = 4 independent experiments with 3 technical replicates/group). (E) Cell survival following 48 h exposure to rotenone with or without verteporfin (n = 3 independent experiments with 3 technical replicates/group). (F) Representative images of endothelial cell marker CD31 (red) and YAP1 (green) in brain tissue from sham and 28 days post-middle cerebral artery occlusion (MCAO) mice, with DAPI staining of nuclei (blue) and merged images highlighting co-localization. n = 4 mice per group; 3 non-consecutive sections per mouse with 5 randomly selected peri-infarct fields analyzed per section. ∗ p < 0.05, ∗∗ p < 0.01, ∗∗∗ p < 0.001, ∗∗∗∗ p < 0.0001. This altered subcellular distribution was further corroborated through confocal microscopy with 3D reconstruction ([103]Fig. 5C). Measurement of the nuclear-to-cytoplasmic intensity ratio of YAP1 fluorescence demonstrated a significant increase in nuclear YAP1 enrichment in PATJ KO cells compared to WT cells. These findings suggest that PATJ may influence YAP1 subcellular localization, potentially affecting its transcriptional activity. The functional effects of pharmacological YAP1 inhibition were tested in HEK293 cells. Treatment with verteporfin, a YAP1 inhibitor, did not affect baseline proliferation. However, verteporfin also further exacerbated the migration deficit observed in PATJ KO cells relative to either WT or untreated KO controls ([104]Fig. 5D), suggesting that YAP1 activity becomes particularly critical for cell motility in the absence of PATJ. Verteporfin also increased metabolic vulnerability to rotenone in PATJ KO cells ([105]Fig. 5E), with PATJ KO cells treated with both verteporfin and rotenone showing reduced survival compared to cells treated with rotenone alone. Together, these findings suggest that the elevated YAP1 activity and nuclear localization in PATJ-deficient cells act as a compensatory mechanism that supports survival and migration under metabolic stress. Inhibiting this compensatory YAP1 activity with verteporfin exacerbates PATJ KO cell vulnerability revealing a functional dependency between PATJ loss and YAP1 activity for cellular survival under conditions of metabolic stress. Notably, verteporfin had a less pronounced effect on WT cells exposed to rotenone, indicating that this YAP1-dependent compensatory mechanism becomes particularly critical in the absence of PATJ. These results establish an important interplay between PATJ and YAP1 in maintaining cellular resilience and motility under pathophysiological stress. To extend these findings to an in vivo context, we examined YAP1 expression and localization in a mouse model of ischemic stroke. Immunohistochemical analysis of brain sections from mice subjected to tMCAO revealed a distinct alteration in YAP1 subcellular localization in endothelial cells 28 days post-stroke compared to sham-operated controls ([106]Fig. 5F). In sham group, endothelial YAP1 expression was primarily localized to the cell surface. In contrast, tMCAO mice exhibited a marked increase in nuclear YAP1 within endothelial cells in the peri-infarct region, as evidenced by the co-localization of YAP1, CD31, and the nuclear marker DAPI. Quantitative analysis of cells positive for all three markers confirmed a significantly higher number of endothelial cells with nuclear YAP1 in tMCAO mice compared to sham controls. These findings suggest that ischemic stroke promotes the nuclear translocation of YAP1 in endothelial cells, potentially as a vascular adaptive response to ischemic injury. 3.4. Temporal profiling of PATJ expression in endothelial cells following stroke To further explore the changes in PATJ after ischemic stroke, we characterized the temporal patterns of PATJ expression following tMCAO in mice ([107]Fig. 6, [108]Fig. 7). We observed a reduction in PATJ expression predominantly within the neuronal cells in the penumbral regions. This decrease was statistically significant when compared to the sham-operated controls and mirrors the decrease in the neuronal marker, NeuN. As the post-stroke timeline progressed, a gradual yet significant increase in endothelial PATJ expression became evident. As the recovery phase progressed, the colocalization of PATJ within the peri-ischemic endothelium gradually increased. By day 28 post-tMCAO, not only did PATJ expression surpass its original levels, but its colocalization with endothelial markers also significantly increased beyond the baseline observed in sham-operated controls. This enhanced colocalization in the later stages post-stroke signifies a possible adaptive augmentation of PATJ in post-stroke endothelial response to drive vascular remodeling and repair. Fig. 6. [109]Fig. 6 [110]Open in a new tab Temporal expression of PATJ in endothelial cells post-tMCAO and colocalization with YAP1. (A)Schematic representation of the peri-ischemic region observed in the tMCAO mouse model. (B) Representative immunofluorescence images showing double staining of Patj (red) and CD31 (green) in brain sections from sham and tMCAO-treated mice at days 1, 3, 7, 14, and 28 post-stroke. (C) Quantitative analysis of Patj/CD31 double positive cells per mm^2 in the peri-ischemic region over time (D) Quantitative analysis of Patj positive cells in the peri-ischemic area over time. (E) Quantitative analysis of CD31 positive signal dynamics in the peri-ischemic zone. n = 4 mice per group; 3 non-consecutive sections per mouse with 5 randomly selected peri-infarct fields analyzed per section. ∗∗p < 0.01, ∗∗∗∗p < 0.0001. (F) Triple immunofluorescence staining showing CD31 (red), YAP1 (green), and PATJ (blue) in brain sections from sham and 7 days post-tMCAO mice. The merged images demonstrate colocalization of all three markers, with white arrowheads indicating triple-positive regions in the vascular structures. (G) Quantification shows the percentage of CD31+/YAP1+/PATJ + triple-positive area in sham versus tMCAO 7d tissues. n = 4 mice per group; 3 non-consecutive sections per mouse with 5 randomly selected peri-infarct fields analyzed per section. ∗∗∗∗p < 0.0001. Fig. 7. [111]Fig. 7 [112]Open in a new tab Temporal shifts in Patj expression in neurons after tMCAO. (A) Representative immunofluorescence images showing double staining of Patj (red) and the neuronal marker NeuN (green) in brain sections from sham and tMCAO mice at 1, 3, 7, 14, and 28 days post-stroke. (B) Quantitative analysis of Patj/NeuN double-positive cells per mm^2 in the peri-ischemic region across the specified time points. (C) Quantitative analysis of Patj positive signal in the peri-ischemic region over time (D) Quantitative analysis of NeuN positive cells in the peri-ischemic zone at each time point post-tMCAO. n = 4 mice per group; 3 non-consecutive sections per mouse with 5 randomly selected peri-infarct fields analyzed per section. ∗p < 0.05, ∗∗p < 0.01, or ∗∗∗p < 0.001, ∗∗∗∗p < 0.0001. To further investigate the spatial relationship between PATJ and YAP1 in the vascular remodeling response, we performed triple immunofluorescence staining for CD31, YAP1, and PATJ in the peri-infarct region at 7 days post-tMCAO ([113]Fig. 6F–G). Confocal microscopy revealed triple colocalization of these proteins within regions of the remodeling vasculature, with quantitative analysis confirming significantly higher triple-positive area in tMCAO samples compared to sham controls. Within CD31-positive vessels, we observed that areas showing elevated YAP1 expression frequently also exhibited increased PATJ expression. These observations suggest a coordinated expression and potential interaction of YAP1 and PATJ during post-stroke vascular remodeling. 3.5. Transcriptomic profiling reveals PATJ's influence on vascular and cellular processes We conducted RNA-sequencing on PATJ KO, PATJ KO with PATJ overexpression (KO + Ov), and WT cell lines, revealing distinct transcriptional landscapes. Principal component analysis (PCA) validated our approach, showing clear group segregation ([114]Supplementary Fig. 3A). A core set of 112 genes were differentially regulated across the three cell types, suggesting essential elements of the PATJ-dependent transcriptional program ([115]Supplementary Fig. 3B). Runt-related transcription factor 1 (RUNX1) is critically involved in hematopoiesis and has been implicated in endothelial functions [[116]29], suggesting a potential link between PATJ modulation and vascular remodeling, a process vital in post-stroke recovery. Hairy/enhancer-of-split related with YRPW motif 1 (HEY1), a transcriptional repressor regulated by the Notch signaling pathway, is known to influence vascular development and integrity [[117]30], aligning with the observed changes in PATJ-mediated gene expression and pointing toward a role in angiogenesis and possibly neurogenesis. Nuclear protein 1 (NUPR1), involved in stress responses and cellular adaptation [[118]31], may offer insights into the cellular mechanisms by which PATJ contributes to the resilience and migration of cells in response to ischemic conditions induced by stroke. Furthermore, Hexokinase 2 (HK2), central to the glycolytic pathway, demonstrated markedly reduced expression levels in PATJ KO cells suggesting a greater dependence on oxidative phosphorylation and supporting the Seahorse assay results also showing a shift in metabolic requirements during cellular stress [[119]32]. Heatmap and volcano plot analyses further highlighted significant upregulation and downregulation of genes associated with immune response, growth regulation, and signal transduction, indicating PATJ's diverse regulatory impact ([120]Fig. 8A–D). Gene Ontology (GO) and KEGG pathway analyses connected the DEGs to crucial stroke-related processes in the KO + Ov group, including vascular development and response to hypoxia, with pathways like 'p53 signaling' and 'Notch signaling' also being enriched ([121]Fig. 8E–F). Fig. 8. [122]Fig. 8 [123]Open in a new tab Transcriptomic landscape and pathway analysis in PATJ-modulated HEK293 cells. (A,B,C) Volcano plots detailing significant gene expression changes for KO + Ov vs. KO, KO vs. WT, and KO + Ov vs. WT, with upregulated genes in red and downregulated genes in blue. (D) Heatmaps of gene expression contrasts for KO + Ov vs. KO vs. WT, with color intensity denoting gene expression levels. (E) Gene Ontology (GO) enrichment analysis bubble chart for the KO + Ov vs. KO comparison. (F) Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway enrichment analysis bubble chart for the KO + Ov vs. KO comparison. 4. Discussion Ischemic stroke results in significant vascular and neuronal damage which initiates a cascade of immune-mediated secondary neuronal injuries in the days and months following stroke. In particular, damage to cerebral endothelial cells (ECs) increases vascular permeability and disrupts the blood-brain barrier which can further exacerbate stroke injury and impair neurorecovery [[124][33], [125][34], [126][35]]. Thus, optimizing the beneficial effects of ECs during stroke recovery is an attractive target to improve stroke outcome [[127]36,[128]37]. PATJ has well established roles in tight junction assembly and maintenance of epithelial cell polarity [[129][8], [130][9], [131][10], [132][11],[133][38], [134][39], [135][40], [136][41], [137][42]]. In our study, we extend recent findings linking PATJ to angiogenesis and provide additional evidence for PATJ involvement in cellular responses to metabolic stress in cells and C. elegans and the physiological response to stroke in vivo in mouse ECs [[138]16]. The establishment of a PATJ KO in HEK293 cell lines, which revealed roles for PATJ in cell proliferation, cell migration, response to mitochondrial stress, and the regulatory dynamics with YAP1. Our findings reveal a significant reduction in cellular migration in PATJ KO HEK293 cells, which can be rescued by inducing the overexpression of PATJ. This aligns with previous studies highlighting PATJ's involvement in regulating directional migration of mammalian epithelial cells and interaction with cytoskeletal proteins [[139]9,[140]28]. Our data also reveal an enhanced vulnerability to mitochondrial stress in PATJ-deficient cells with reduced proliferation in rotenone and altered oxygen consumption following oligomycin, highlighting PATJ's contribution to cellular resilience against metabolic challenges. This aspect is particularly relevant in the context of stroke, where cellular energy demands are acutely altered. The ability of cells to withstand and repair tissues following ischemia is critical for neurorecovery, suggesting that interventions aimed at modulating PATJ expression or its targets could improve cellular resilience in ischemic injury. The regulatory role of PATJ in endothelial cells under hypoxic conditions is further supported by our bioinformatics analyses of multiple existing HUVEC RNA-seq datasets. We observed significant upregulation of PATJ in endothelial cells following hypoxia exposure, consistent with the increased PATJ expression seen in endothelial cells during the recovery phase post-stroke in our mouse model. The strong correlation between PATJ and key angiogenic factors such as VEGFA, VEGFC, and PECAM1 in HUVECs further suggests PATJ's involvement in angiogenic processes. The regulatory interplay between PATJ and YAP1, especially under conditions of PATJ depletion, introduces a new angle for further study in understanding EC response mechanisms to ischemic stroke, especially considering YAP1 is highly enriched in brain vascular cells (see [141]Supplementary Fig. 1; [[142]26]). The observed changes suggest that PATJ deletion may trigger YAP1 signaling compensatory mechanisms. Emerging evidence suggests that nuclear translocation of the transcription factor, YAP1 modulates the expression of genes involved in angiogenesis [[143][43], [144][44], [145][45], [146][46], [147][47], [148][48], [149][49], [150][50], [151][51], [152][52], [153][53], [154][54], [155][55]]. Intriguingly, our in vivo studies corroborate and extend these findings. We observed a marked increase in nuclear YAP1 within endothelial cells in the peri-infarct region, suggesting a potential vascular response to ischemic injury. While the precise role of YAP1 in post-stroke angiogenesis remains to be fully elucidated, this nuclear translocation may indicate its involvement in vascular adaptation processes. Our functional experiments with verteporfin, a specific inhibitor of YAP1 activity, provide evidence for a compensatory role of YAP1 signaling in PATJ-deficient cells under metabolic stress conditions. Inhibition of YAP1 reduced cell survival in PATJ KO cells exposed to rotenone-induced stress suggesting that the increased nuclear YAP1 localization observed in these cells may represent an adaptive response to preserve cellular resilience. These observations align with emerging evidence suggesting that YAP1 can function as a stress-responsive factor that promotes cellular adaptation and survival under various stress conditions [[156][56], [157][57], [158][58], [159][59]]. In the context of PATJ deficiency, our data indicates that YAP1 activation may represent a critical backup mechanism that becomes essential for cellular resilience when PATJ-dependent pathways are compromised. This compensatory relationship may have important implications for understanding how cells adapt to ischemic injury and for developing therapeutic strategies that preserve or enhance these adaptive responses. Angiogenesis, which principally involves ECs, begins in the zones surrounding the ischemic event within the first 12 h and can persist for more than three weeks [[160]60,[161]61]. Angiogenesis is necessary for the repair and remodeling of vascular structures and also promotes neural plasticity, as evidenced in both clinical and preclinical stroke studies [[162]62]. The increased nuclear localization of YAP1 in endothelial cells of the peri-infarct region could represent part of a complex adaptive response, potentially influencing vascular remodeling in the recovering brain tissue. Thus, further understanding the mechanistic connections between PATJ and YAP1 may aid in uncovering new therapeutics to enhance post-stroke angiogenesis and tissue repair. The dynamic regulation of PATJ expression in ECs following ischemic stroke, characterized by an initial decrease and subsequent increase, suggests a tightly controlled response to ischemic injury. This biphasic pattern might reflect the role of PATJ in interacting with EC repair processes post-stroke. Interestingly, our findings in C. elegans demonstrate that the specific upregulation of MPZ-1, the C. elegans ortholog of PATJ, under hypoxic conditions is particularly relevant to understanding PATJ's potential role in ischemic stroke adaptation. As ischemic stroke is characterized by reduced blood flow leading to hypoxia and metabolic stress in the affected brain regions, the observation that PATJ expression is induced by hypoxia in C. elegans suggests that this response may be an evolutionarily conserved mechanism to cope with oxygen deprivation. Our findings in the C. elegans model provide further evidence for the evolutionarily conserved role of PATJ in conferring neuronal resilience against hypoxic injury. Knockdown of the PATJ homolog mpz-1 resulted in abnormal neuronal morphology, an effect that was exacerbated when combined with hypoxia exposure in adult worms. Moreover, we observed a significant increase in mortality among mpz-1 deficient worms subjected to hypoxia, suggesting that PATJ/MPZ-1 is not only crucial for maintaining neuronal structural integrity but also for promoting overall organismal survival under hypoxic stress. This supports our mouse data showing altered PATJ expression in neurons following ischemic stroke and suggests that PATJ is critical for maintaining neuronal health and counteracting damage caused by oxygen deprivation. The hypoxia-sensitive phenotype and increased mortality in mpz-1 deficient worms parallel the increased vulnerability of PATJ KO mammalian cells to mitochondrial stressors, emphasizing its function in adapting to metabolic challenges at both the cellular and organismal levels. Given that ischemic stroke triggers hypoxia and metabolic dysfunction, the evolutionarily conserved ability of PATJ to orchestrate adaptive responses may have important implications for understanding its influence on stroke outcomes. Future work will be necessary to understand whether modulating PATJ expression also modulates vascular remodeling to enhance stroke recovery. Conversely, the sustained decrease in neuronal PATJ expression following stroke further emphasizes the need for additional research in understanding the cell-type-specific regulation of PATJ to enhance stroke recovery. Our transcriptomic analysis, centered on PATJ's modulation, illuminates its pivotal role in regulating genes and pathways critical for vascular remodeling and recovery post-stroke. Notably, the downregulation of RUNX1 and HEY1 in PATJ KO cells, implicated in endothelial functions and angiogenesis [[163]29,[164]30,[165]63], underscores PATJ's influence on vascular development. This, coupled with NUPR1's involvement in cellular stress responses and HK2's role in metabolic adaptation [[166]31,[167]32], highlights metabolic influences regulated by PATJ. Notably, the upregulation of noggin (NOG) with PATJ overexpression and its down-regulation in the PATJ KO, suggests a previously unexplored connection between PATJ and this key regulator of neural development and neurogenesis [[168]64]. Similarly, RBM15 stands out with its substantial upregulation in the KO + Ov cell line and downregulation in the PATJ KO. RBM15 is integral to RNA metabolic processes, influencing not only mRNA splicing but also chromatin organization and gene expression of pathways critical for cellular identity and function. RBM15 dysregulation along with enrichment of the Notch signaling pathway in KEGG analysis aligns with our findings implicating PATJ in the modulation of vascular-related functions [[169]65]. Interestingly, the chemokine C-X-C motif receptor (CXCR)-4- was also downregulated in the PATJ KO cell line, and is essential for chemotaxis and hematopoietic cell lineages [[170]66] as well as cooperates with Notch signaling to facilitate arterial differentiation [[171]63]. Taken together, these findings support a role for PATJ in influencing cell migration and differentiation. In contrast, genes such as CHAC1 (involved in glutathione metabolism and apoptosis), SLC6A9 (a neurotransmitter transporter), and INHBE (a member of the TGF-beta superfamily) exhibited considerable downregulation in KO + Ov cells and upregulation in PATJ KO cells. These changes may reflect PATJ's involvement in the regulation of redox homeostasis, neurotransmission, and cell growth signaling pathways. The deregulation of these genes with genetic manipulation of PATJ expression coupled with the metabolic and cell biological phenotypes demonstrated in this study could suggest possible mechanistic targets for PATJ to exert effects on cellular stress responses, synaptic signaling, and tissue homeostasis. These findings, aligned with the validated GO and KEGG pathway functions, suggest PATJ's integral role in orchestrating cellular resilience, angiogenesis, and neurogenesis through pathways like Notch signaling. Our transcriptomic profiling in HEK293 cells suggests that PATJ may influence multiple molecular pathways pertinent to angiogenesis, metabolic regulation, and cellular stress responses. Though further research is required to understand the relevance of PATJ to these pathways following stroke in vivo, this is a valuable starting point for future investigations examining how PATJ might interface with cellular pathways relevant to ischemic remodeling and neurovascular repair. The significant colocalization of PATJ and YAP1 in endothelial cells post-stroke provides new insights into potential molecular interactions during vascular remodeling. Our triple immunofluorescence analysis demonstrated that these proteins colocalize within CD31-positive vessels in the peri-infarct region, specifically during the active angiogenic phase. This observation, combined with our in vitro evidence showing altered YAP1 localization in PATJ KO cells, points to a potential relationship between PATJ and YAP1 in the vascular response to ischemic injury. While the precise functional relationship between these proteins requires further investigation, their coordinated expression in remodeling vessels highlights a potentially important molecular interaction that may influence angiogenic responses after stroke. However, we acknowledge that additional pathways may converge on YAP1 during stroke-induced stress. Future studies employing conditional knockout models in endothelial cells—or advanced imaging techniques to monitor YAP1 subcellular dynamics in real time—could help distinguish whether loss of PATJ alone drives YAP1 nuclear translocation or if other ischemia-related signals also modulate YAP1. In particular, proximity ligation or co-immunoprecipitation assays might clarify whether PATJ physically binds YAP1 to govern its localization, or alternatively, acts through intermediary effectors. These directions represent a critical next step in delineating how PATJ deficiency intersects with YAP1's regulatory network under the complex conditions of stroke recovery. Extending the foundational work by Medina-Dols et al. on PATJ's role in modulating endothelial to mesenchymal transition (EndMT) through key pathways [[172]16], our investigation provides additional insights and corroborates key findings with complementary techniques. Employing CRISPR-Cas9 technology, we illuminate PATJ's broader influence on cellular dynamics and vascular remodeling. Our findings reveal significant PATJ-driven alterations in cellular migration and mitochondrial stress response, highlighting a regulatory interplay with the YAP1 signaling pathway, particularly highlighting YAP1's nuclear translocation—a phenomenon also partially supported by Medina-Dols and colleagues findings. Our RNA-seq analysis reveals extensive transcriptional reprogramming under PATJ modulation, identifying dysregulation in genes critical to vascular development, stress response, and metabolism. Building upon Medina-Dols et al.’s observations of sustained PATJ downregulation [[173]16], our in vivo analysis showcases a dynamic, time-dependent expression pattern in endothelial cells, suggesting PATJ's integral role in both the immediate response and subsequent recovery processes. This biphasic expression underscores the complexity of PATJ's involvement beyond EndMT, implicating it in essential pathways for vascular development and recovery. Thus, this work significantly extends our understanding of PATJ's role in stroke pathophysiology and supports further work in this pathway to pave new avenues for targeted therapeutic interventions in stroke rehabilitation and vascular repair. It is important to acknowledge the inherent limitations of the tMCAO model used in this study. Despite standardization efforts, biological variability in individual mice may influence the extent of ischemic injury and subsequent vascular responses. Additionally, while the tMCAO model is widely accepted for studying ischemic stroke, it does not fully recapitulate all aspects of human stroke pathophysiology, particularly regarding comorbidities commonly present in stroke patients. These limitations should be considered when translating our findings to clinical applications. In conclusion, our research advances the understanding of PATJ's multifaceted roles in cellular and vascular physiology, particularly its evolutionarily conserved function in modulating cellular responses to hypoxic stress, and highlights the potential of targeting PATJ-mediated molecular network for therapeutic intervention in stroke recovery. Future studies will explore the detailed mechanisms by which PATJ influences EC function and vascular remodeling in stroke recovery and other pathologic states, further illuminating its role in disease modulation and therapeutic potential. CRediT authorship contribution statement Mengqi Zhang: Writing – original draft, Visualization, Validation, Methodology, Investigation, Formal analysis, Data curation. Wei I. Jiang: Visualization, Validation, Methodology, Investigation, Formal analysis. Kajsa Arkelius: Validation, Methodology, Investigation. Raymond A. Swanson: Writing – review & editing, Supervision, Project administration. Dengke K. Ma: Writing – review & editing, Supervision, Project administration, Investigation, Data curation. Neel S. Singhal: Writing – review & editing, Supervision, Resources, Investigation, Funding acquisition, Conceptualization. Funding acknowledgments