Abstract Endothelial barrier dysfunction and the resulting vascular injury are responsible for multiorgan failure in sepsis. Myeloid C-type lectin domain family 5 member A (CLEC5A) is a pattern recognition receptor involved in host defense against infection. Mice lacking CLEC5A were resistant to cecal ligation and puncture (CLP)–induced polymicrobial sepsis and lipopolysaccharide (LPS)–induced endotoxemia, as observed by decreased mortality. Single-cell RNA sequencing revealed transcriptomic heterogeneity of vascular endothelial cells in CLEC5A-deficient lungs following CLP. Endothelial-specific knockdown of CLEC5A improved survival of CLP-challenged mice, which was completely ineffective with reexpression of endothelial CLEC5A. The survival benefits were attributed to alleviated inflammatory storm and vascular leakage. Furthermore, endothelial CLEC5A deficiency protected mice against Escherichia coli–induced pneumonia. In vitro, CLEC5A deletion maintained trans-endothelial electrical resistance, and inhibited adhesion and trans-endothelial migration of monocytes/neutrophils under LPS stimulation. The study unveils the importance of CLEC5A in regulating endothelial barrier function and suggests endothelial CLEC5A as a therapeutic target for pneumonia or sepsis-causing bacterial infection. __________________________________________________________________ Endothelial CLEC5A deficiency protects against inflammatory lung injury and provides survival benefits in sepsis. INTRODUCTION Sepsis is a life-threatening condition with multiorgan failure caused by a dysregulated host response to infection ([48]1, [49]2). For many of the patients who survive, it is associated with a long-term morbidity. Although optimal management is fundamental in the intensive care setting, early recognition and rapid intervention during sepsis are required to improve outcomes ([50]3). Ongoing research continues to increase understanding of molecular and biochemical profiles of patient populations and improves disease definition, diagnosis, and pharmacological strategies to achieve more favorable outcomes in sepsis ([51]4). Pulmonary infection is a common cause of sepsis ([52]4). Patients with sepsis are at a high risk of developing acute respiratory distress syndrome (ARDS), which occurs in up to 50% of patients ([53]5). Lung injury is one of the major attributable cause of death in sepsis. It is characterized by alveolar-capillary barrier dysfunction and resulting vascular leakage and subsequent influx of albumin-rich edema fluid into the alveolar cavity, causing hypoxemia ([54]6). Over several decades, the incidence of sepsis has risen and, in sepsis-associated organ dysfunction, especially ARDS, morbidity and mortality remain high. The necessity for advances in our understanding of the pathophysiological mechanisms and development of effective therapeutic strategies cannot be overstated. Sepsis and pneumonia share several similar clinical presentations and disease characteristics, in which infections and inflammation are the key. In response to infection, immunity maintains homeostasis between immunopathology and immunosuppression. A failure of immune homeostasis is considered the key pathophysiology of sepsis ([55]2). Activation of the local innate immune response is important for host defense against pathogen spreading. This activation may become progressive during sepsis, causing tissue injury and leading to multiorgan failure and ultimately death ([56]7). Although the major cell sources for inflammatory cytokines in sepsis are myeloid cells, nonmyeloid cells such as endothelial cells are also involved and have been highly valued ([57]8). Endothelial cells are an unconventional type of immune cells that lines blood vessels and participates in immune and inflammatory responses upon activation ([58]9). In sepsis, endothelial cells exhibit phenotypic changes toward pro-inflammation, pro-adhesion, pro-coagulation, and pro-apoptosis, and these changes contribute to leukocyte recruitment and pathogen elimination. However, severe and/or persistent activation causes damage to the endothelium, resulting in barrier dysfunction and exacerbating local inflammation and tissue damage ([59]10). The endothelium is considered a target in the treatment of sepsis and limiting barrier dysfunction is a promising avenue for preventing the development of lung injury. Myeloid C-type lectin domain family 5 member A (CLEC5A) is a pattern recognition receptor (PRR) and plays a critical role in innate immunity against viral infection ([60]11, [61]12). Apart from being a promiscuous PRR to viruses, CLEC5A has recently been noted in host defense against bacterial infection ([62]13, [63]14). Sung et al. reported that deletion of CLEC5A or administration of anti-CLEC5A monoclonal antibodies alleviated lung injury and improved survival of Pseudomonas aeruginosa–challenged mice ([64]14). These observations imply the therapeutic potential of targeting CLEC5A in infectious disorders and the resulting lung injury; however, the underlying cellular and molecular mechanism has not been fully unveiled. Current studies mostly focus on CLEC5A in myeloid cells, and little is known about its function in other cell types. A comprehensive understanding of CLEC5A is needed to further its implications as a therapeutic target in the clinical setting. In the present study, we uncover the role of CLEC5A in endothelial cells regarding phenotypic modifications and functional alterations during inflammation and bacterial infection, as well as in cellular interaction with immune cells. The immune response in the local environment of sepsis was investigated by modeling the interaction of endothelial cells and immune cells under lipopolysaccharide (LPS) stimulation in vitro. Deficiency of endothelial CLEC5A prevented against barrier failure and lung injury and improved survival in murine sepsis models. This study broadens the profile of CLEC5A to the endothelium and improves the existing therapeutic algorithms of CLEC5A in lung injury. We suggest that CLEC5A could be a target in translational endothelial-driven therapeutic approaches to bacterial pneumonia and sepsis. RESULTS Global deletion of CLEC5A improves the survival of mice with endotoxemia or polymicrobial sepsis To explore the potential targets of gene therapy for sepsis and sepsis-induced lung injury, we obtained and analyzed the gene expression profiles of four datasets ([65]GSE28750, [66]GSE206635, [67]GSE1871, and [68]GSE34901) from the NCBI Gene Expression Omnibus (GEO) database ([69]https://ncbi.nlm.nih.gov/geo/) ([70]Fig. 1A). The datasets [71]GSE28750 collect data from blood samples from patients with sepsis and healthy volunteers, and data of [72]GSE206635 are from the lungs of patients with COVID-19. [73]GSE1871 and [74]GSE34901 include data from lungs from LPS-induced murine lung injury and Escherichia coli–induced murine pneumonia, respectively. The differentially expressed genes (DEGs) were derived from the four datasets with the linear models for microarray data (limma) following the screening criteria of fold change (FC) > 1.5 and P < 0.05. A total of 154 robust genes were screened out according to the robust rank aggregation with a score of < 0.05 and frequency ≥ 3 (number of datasets). Among genes in C-type lectin domain family, CLEC5A is top ranked and significantly up-regulated in all four datasets (Score < 0.0001 and frequency = 4) ([75]Fig. 1B). CLEC5A is widely expressed in human tissues and highly expressed in the lungs based on the Genotype-Tissues Expression (GTEx) dataset and the FANTOM5 dataset (fig. S1), which were obtained from an open source database, the Human Protein Atlas (HPA; [76]http://www.proteinatlas.org). Fig. 1. Genetic knockout of CLEC5A improves the survival of mice with endotoxemia or polymicrobial sepsis. [77]Fig. 1. [78]Open in a new tab (A) Analysis of the gene expression profiles of datasets ([79]GSE28750, [80]GSE206635, [81]GSE1871, and [82]GSE34901) downloaded from the NCBI GEO database and identification of CLEC5A as a potential gene related to sepsis and sepsis-induced acute lung injury. Top DEGs among C-type lectin domain family were displayed based on the limma with P < 0.05 and FC > 1.5 and sorted by the robust rank aggregation. Genes on white panel meet score (P value) < 0.05 and frequency (Freq) ≥ 3. Blue font highlights the top-ranked gene. NA, not applicable. (B) mRNA expression of CLEC5A in blood samples from patients with sepsis ([83]GSE28750), lung tissues from patients with COVID-19 ([84]GSE206635), and lungs from mice with LPS-induced lung injury ([85]GSE1871) and E. coli–induced pneumonia ([86]GSE34901). (C) Two mouse models of sepsis were established in this study, including bacterial LPS-induced endotoxemia and CLP-induced polymicrobial sepsis. PMVECs were isolated from corresponding mice. (D) Representative images of immunofluorescence staining for CLEC5A with DAPI counterstaining of nuclei in lung tissues, 4 hours after LPS or 12 hours after CLP (biological replicates, n = 6 per group). Scale bars, 100 μm. (E) Western blotting analysis and representative blots showing the expression of CLEC5A in PMVECs from mice challenged by LPS or CLP (biological replicates, n = 6 per group). h, hours. (F) mRNA expression of CLEC5A in mouse PMVECs (biological replicates, n = 6 per group). (G) Schematic diagram of CLEC5A genetic knockout (KO) mice (CLEC5A^−/−) generation. CLEC5A^−/− mice were generated on a C57BL/6 background by the CRISPR-Cas9 system, and WT littermates were used as the control. (H) Survival rate of mice after LPS or CLP challenge (biological replicates, n = 10 per group). The statistical significance was determined by repeated measures two-way ANOVA analysis using the Geisser-Greenhouse correction. The statistical significance between survival curves was determined by P value using the log-rank (Mantel-Cox) test. n.s., not significant. By analyzing these public datasets, we hypothesize that high expression of CLEC5A might play a role in the pathophysiology of these lung diseases caused by bacterial or viral infection, in which endothelial barrier dysfunction is a key. The present study aims to investigate the role of CLEC5A in sepsis-associated lung injury and endothelial barrier dysfunction using two models of sepsis, intraperitoneal injection of LPS and cecal ligation and puncture (CLP) surgery ([87]Fig. 1C). The expression of CLEC5A in the lungs was up-regulated upon LPS or CLP challenge (fig. S2, A and B). During the induction of experimental sepsis, CLEC5A exhibited differential expression at different time points (4, 12, 24, and 72 hours) post-LPS or CLP but maintained an up-regulated trend in general. Immunofluorescence staining showed an increase in CLEC5A expression in lung tissues 4 hours after LPS or 12 hours after CLP challenge ([88]Fig. 1D). Of note, a significant increase in the expression of CLEC5A was found in primary pulmonary microvascular endothelial cells (PMVECs) isolated from septic mice following LPS or CLP induction for 4 or 12 hours ([89]Fig. 1, C, E, and F). CLEC5A genetic knockout (CLEC5A^−/−) mice were used in this study ([90]Fig. 1G), and the efficacy of CLEC5A deletion was validated in different tissues, including the lung, brain, heart, kidney and liver, from both male and female CLEC5A^−/− mice (fig. S3). Global deletion of CLEC5A significantly increased survival of mice in the setting of LPS-induced endotoxemia and CLP-induced polymicrobial sepsis ([91]Fig. 1H). All wild-type (WT) littermates died within 108 hours after LPS challenge, whereas only 50% of CLEC5A^−/− mice died within the same time period and 40% survived until 156 hours post-LPS. In the model of polymicrobial sepsis, 80% of CLEC5A^−/− mice survived until 72 hours post-CLP compared to 100% death during the same timeframe in WT littermates. Depletion of CLEC5A reduced mortality from 100 to 60% in mice within 156 hours following CLP or LPS challenge. Improvement in the survival of CLEC5A^−/− mice challenged by CLP or LPS suggests that CLEC5A is a crucial factor in the development or progression of sepsis. scRNA-seq profiles of major cell types in the lungs following CLP: Focusing on endothelial cells To further explore the role of CLEC5A in sepsis-induced lung injury, a comprehensive single-cell atlas of the lungs was generated by single-cell RNA sequencing (scRNA-seq) ([92]Fig. 2A). Mice were subjected to CLP, and the lungs were collected 12 hours postsurgery. Single-cell suspensions were prepared from lung tissues of three CLEC5A^−/− mice and three WT littermates. Pooled cells from these six mice were subjected to scRNA-seq processing using the 10×Genomics Chromium system. A total of 55,177 single cells, including 29,306 cells from WT lungs and 25,871 cells from CLEC5A^−/− lungs, were obtained after removal of multiplets and filtration of low-quality cells. The software programs Cell Ranger and Seurat were used to analyze the sequencing data and generate the single-cell transcriptome. According to principal components analysis (PCA) and clustering analysis, total cells were grouped into 34 distinct clusters by uniform manifold approximation and projection (UMAP), and these cell clusters were further identified as different cell types based on the expression of their marker genes using the FindAllMarkers of Seurat ([93]Fig. 2B). UMAP analysis showed a total of 14 distinct cell types comprising the mouse lungs, including granulocytes, T cells, B cells, alveolar-macrophages (AM), monocytes, macrophages, natural killer T cells (NKT), mesothelium, natural killer cells (NK), endothelial cells, fibroblasts, dendritic cells (DC), epithelial cells, and smooth muscle cells (SMC). They could be grouped into immune cells (including granulocyte, T cell, B cell, AM, monocyte, macrophage, NKT, NK, and DC) and nonimmune cells (including mesothelium, endothelial, fibroblast, epithelial, and SMC). In general, the septic lungs were largely populated by the immune cells, in which granulocytes, T cells and monocytes were the most prevalent types. Overall, there was an increase in the relative percentage of immune cells in lung tissues of CLEC5A^−/− mice compared to WT lungs following CLP. Among immune cells, the proportion of most cell types increased in CLEC5A-deficient lungs. Notably, the proportion of monocytes was largely decreased in the lungs of CLEC5A-deficient mice in the setting of CLP-induced lung injury. Among nonimmune cells, endothelial and epithelial cells were relatively dominant in the septic lungs of CLEC5A^−/− mice. Fig. 2. scRNA-seq analysis of endothelial cells in the lungs after CLP. [94]Fig. 2. [95]Open in a new tab (A) Overview of the study design for scRNA-seq. (B) UMAP plot of 14 cell types identified from the lungs of CLEC5A^−/− mice and WT littermates (biological replicates, n = 3 per group), including granulocyte, T cell, B cell, AM, monocyte, macrophage, NKT, mesothelium, NK, endothelial, fibroblast, DC, epithelial, and SMC. They were grouped into immune cells and nonimmune cells. (C) UMAP plot of endothelial features showing the distribution of established marker gene PECAM1. Gene expression level is shown by purple color grading. (D) Violin plot depicting the expression level of endothelial cell-type marker gene PECAM1. (E) Representative immunofluorescence images showing coexpression of CLEC5A and CD31 (encoded by gene PECAM1) in lung tissues (biological replicates, n = 6 per group). Scale bars, 50 μm. Percentage of CLEC5A^+CD31^+ in total CD31^+ cells was quantified. (F) Relative expression of PECAM1 mRNA and CD31 protein levels in PMVECs. The statistical significance between two groups was determined by an unpaired two-tailed t test and comparisons among more than two groups by one-way ANOVA analysis. (G) UMAP plot of endothelial cell clusters in CLEC5A^−/− mice (red dots) and WT littermates (blue dots). (H) UMAP plot of endothelial cell subtypes and cell percentage, including endothelial cells of lymphatic, aerocyte, gCap, arterial, and venous types. (I) GO and KEGG enrichment analysis for DEGs in endothelial cells between CLEC5A^−/− mice and WT littermates (FC > 1.5 or FC < 0.67 and P < 0.05). Bar plots of top 10 enriched GO-BP and KEGG pathways. The significance of enrichment was determined by P value < 0.05. Although much is known about CLEC5A in immune cells, there is little information about the role of CLEC5A in endothelial cells. The distribution and expression level of the marker gene PECAM1, which has been used previously for identification of endothelial cells in whole lung tissues, were displayed by UMAP analysis ([96]Fig. 2C). Mice lacking CLEC5A had a slightly decreased trend in the number of endothelial cells ([97]Fig. 2B). Changes in the proportion of endothelial cells were further examined by specific marker gene expression. We found the expression level of PECAM1 was significantly up-regulated in CLEC5A^−/− lungs compared to control lungs ([98]Fig. 2D). As the gene PECAM1 encodes for the endothelial surface marker CD31, we used CD31 staining to identify endothelial cells. Immunofluorescence staining confirmed the expression of CLEC5A in CD31-positive endothelial cells in mouse lung tissues ([99]Fig. 2E). There was a decreased trend of CD31 staining in septic lungs, whereas coexpression of CLEC5A and CD31 was increased after CLP. The mRNA and protein expression of PECAM1/CD31 was significantly down-regulated in the lungs of CLP mice but up-regulated in CLEC5A^−/− mice compared to WT mice after CLP ([100]Fig. 2F). CLEC5A was significantly up-regulated in CD31-positive endothelial cells in the lungs upon septic infection. Characterization of endothelial cell subtypes and transcriptome analysis by scRNA-seq in CLEC5A^−/− mice An in-depth analysis was performed examining aspects of cell subtype characterization, cell proportion, and transcriptomics in the endothelium ([101]Fig. 2G). Within the endothelial cell clusters (a total of 1621 single cells), five subtypes were identified, including lymphatic, aerocyte, general capillary (gCap), arterial, and venous types ([102]Fig. 2H). The cellular composition was further analyzed by calculating the relative percentage of each subtype. The gCap subtype was the most populous of the endothelial cell subtypes, and there was no significant difference in the proportion of each subtype between two genotypes. Generally, endothelial cells belong to two main subsets: lymphatic and vascular. The vascular endothelial cells contain the subtypes arterial, venous, and capillary (aerocyte and gCap). Capillary endothelial cells were the most prevalent type in the endothelium of septic lungs ([103]Fig. 2H). We further examined the expression pattern of CLEC5A in theses endothelial subtypes in the lungs of WT mice after CLP challenge (fig. S4). Immunostaining showed that CLEC5A was highly expressed in all endothelial subtypes, with the use of cell marker Gap Junction Protein Alpha 5 (GJA5) for arterial type, Von Willebrand Factor (VWF) for venous type, Endothelin Receptor Type B (EDNRB) for capillary type, Protein Tyrosine Phosphatase Receptor Type B (PTPRB) for gCap type, and Prospero Homeobox 1 (PROX1) for lymphatic type. The DEGs between the endothelial cells comprising CLEC5A^−/− and WT lungs that meet the criteria of FC >1.5 or <0.67 and P < 0.05 were subjected to Gene Ontology (GO) enrichment analysis and Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway enrichment analysis ([104]Fig. 2I). Among the top 10 enriched GO and KEGG pathways, tumor necrosis factor (TNF) signaling pathway, Staphylococcus aureus infection, interleukin-17 (IL-17) signaling pathway, inflammatory response to wounding, and activation of membrane attack complex are strongly associated with the immune/inflammatory response and the development of sepsis and lung injury ([105]15–[106]17). These DEGs were also significantly enriched in genes involved in negative regulation of cell migration and negative regulation of cell motility, which are involved in immune cell recruitment and thus immune/inflammation regulation ([107]18). CLEC5A^−/− transgenic mice are resistant to sepsis-induced lung injury and inflammation There were significant abnormalities in the lungs following CLP ([108]Fig. 3). The bronchoalveolar lavage fluid (BALF) and lungs were harvested 12 hours post-CLP. The morphological changes in the lungs were attenuated in CLEC5A^−/− mice following CLP, as evidenced by less pulmonary hemorrhage and edema ([109]Fig. 3A), as well as decreased lung index (lung-to-body weight ratio) and lung wet-to-dry weight ratio ([110]Fig. 3, B and C). Histopathological analysis was carried out with hematoxylin and eosin (H&E) staining ([111]Fig. 3D) and lung injury score ([112]Fig. 3E). H&E staining displayed alveolar hemorrhage, alveolar space enlargement, and inflammatory cell infiltration in the lungs of mice subjected to CLP-induced polymicrobial sepsis, which correspond to a higher lung injury score. The severity of CLP-induced lung injury was significantly lessened in CLEC5A^−/− mice compared to WT littermates. CLP induced pulmonary fluid collections, and the transudate sampled by BALF was enriched in infiltrated cells, including leukocytes, neutrophils, and monocytes, in comparison with infiltrate-poor cells in mice that received sham operation ([113]Fig. 3F). Along with inflammatory cell infiltration, the levels of pro-inflammatory cytokines including TNF-α and IL-6, and chemokines including monocyte chemotactic protein 1 (MCP-1) and C-X-C motif chemokine ligand 5 (CXCL5), were increased in the lungs from mice challenged by CLP ([114]Fig. 3G), as well as in the BALF ([115]Fig. 3H). Deletion of CLEC5A inhibited inflammatory cell influx into BALF and decreased levels of pro-inflammatory cytokines and chemokines in the BALF and lungs of CLP-challenged mice ([116]Fig. 3, G and H). No significant difference was observed between CLEC5A^−/− and WT mice under sham conditions, indicating that dysregulated CLEC5A might contribute to lung injury only under pathological conditions. Fig. 3. Deletion of CLEC5A protects against CLP-induced lung injury in mice. [117]Fig. 3. [118]Open in a new tab CLEC5A^−/− transgenic mice and WT littermates were subjected to CLP to induce polymicrobial sepsis, and mice received sham operation as the control. The lungs were collected at 12 hours post-CLP. (A) Gross morphology of the lung at 12 hours post-CLP operation. (B) Lung index shown as the percentage of lung weight to body weight. (C) Ratio of lung wet-to-dry weight. (D) Representative images of H&E staining of lung tissues in sham and CLP mice. In the CLP group, images on the right panel represent the region marked by a black square from the left panel. Scale bars, 100 μm. (E) Severity of CLP-induced lung injury determined by inflammation, edema, hemorrhage, and alveolar septal thickening, corresponding to H&E staining. (F) Inflammatory cell counts in BALF, including total leukocytes and differential neutrophils and monocytes. (G and H) Levels of inflammatory cytokines and chemokines in BALF and lung tissues, including TNF-α, IL-6, MCP-1, and CXCL5. The statistical significance for lung wet/dry weight ratio and lung injury score was determined by the Kruskal-Wallis test, and others were determined by one-way ANOVA analysis (biological replicates, n = 6 per group for each experiment). Endothelial knockdown of CLEC5A improves survival of mice with polymicrobial sepsis CLP-induced lung injury is primarily due to alveolar-capillary barrier dysfunction. The results of scRNA-seq transcriptome analysis in CLEC5A-deficient endothelium suggest the importance of endothelial CLEC5A in the development of lung injury and therefore mortality in sepsis. Endothelial specific expression or knockdown of CLEC5A was used in CLEC5A^−/− and WT mice under the endothelial-specific promoter Tie1 by adeno-associated virus type 9 (AAV9) system ([119]Fig. 4A). We confirmed the cellular target of in vivo AAV9-Tie1 infection in the lungs of CLEC5A^−/− mice. Five subtypes (arterial, venous, gCap, aerocyte, and lymphatic) of endothelial cells were identified from scRNA-seq. Immunostaining displayed that the expression of CLEC5A was notably increased in each subtype of endothelial cells via the AAV9-Tie1 system in CLEC5A^−/− mice (fig. S5, A and B). Primary PMVECs lining the lung’s small blood vessels, including capillaries, arterioles and venules, were isolated from WT mice. The efficiency of AAV9 carrying Tie1 promoter approach to endothelial-specific CLEC5A overexpression (CLEC5A^Tie1-oe) or knockdown (CLEC5A^Tie1-sh) was validated in PMVECs (fig. S6A). Alveolar epithelial cells (AECs) are an important cell type in the lung and responsible for alveolar-capillary barrier function, which are closely associated with pulmonary edema and inflammation. In nonendothelial AECs, CLEC5A expression was not affected by the AAV9-Tie1 system (fig. S6B). These data indicated the efficacy of promoter Tie1 to selectively target CLEC5A in pulmonary endothelial cells via AAV9 in vivo infection, whereas no specificity in endothelial cell subtypes. The restoration of CLEC5A expression level was examined in PMVECs of CLEC5A^−/− mice after AAV9-Tie1 injection ([120]Fig. 4B). We found that the level of endothelial CLEC5A in CLEC5A^−/− mice rescued by AAV9-Tie1 (CLEC5A^Tie1-oe) infection was still lower than its endogenous level in WT mice. Mice with endothelial CLEC5A overexpression showed no significant difference in survival rate after CLP challenge compared with WT mice, 100% of which died within 72 hours postoperation ([121]Fig. 4C). In contrast, the survival after CLP was significantly improved in mice lacking endothelial CLEC5A. Thirty percent of CLEC5A^Tie1-sh mice survived until the end of study. In the setting of global deletion of CLEC5A, 50% of mice survived until 156 hours post-CLP, whereas restoration of endothelial CLEC5A expression reduced the survival rate to 20% in CLEC5A^−/− mice ([122]Fig. 4D), which is still higher than WT mice. This survival difference could be attributed to the level of endothelial CLEC5A, which was not fully restored by the AAV9-Tie1 system. We suggest endothelial CLEC5A as a potential pathological factor in CLP-induced polymicorbial sepsis, and its level may be correlated with the survival. Whether the magnitude of CLEC5A expression within different subtypes of endothelial cell influences the survival outcomes and the influence of each subtype needs more studies. Fig. 4. Specific deletion of endothelial CLEC5A improves the survival after CLP and decreases pulmonary endothelial barrier permeability. [123]Fig. 4. [124]Open in a new tab (A) In vivo endothelial-specific expression or knockdown of CLEC5A was carried out via tail vein injection of AAV9 under promoter Tie1. (B) Expression of CLEC5A in PMVECs isolated from WT and CLEC5A^−/− mice (biological replicates, n = 4 per group). (C) Survival rate of WT mice with endothelial-specific CLEC5A overexpression or knockdown (CLEC5A^Tie1-oe or CLEC5A^Tie1-sh) (biological replicates, n = 10 per group). (D) Survival rate of CLEC5A^−/− mice with CLEC5A^Tie1-oe (biological replicates, n = 10 per group). The statistical significance between survival curves was determined by P value using the log-rank test for trend. (E) Experimental design for in vivo and in vitro studies of pulmonary endothelial barrier function. (F) Pulmonary microvascular albumin leakage was determined as the content of EB dye-labeled albumin in the lung (biological replicates, n = 6 per group). (G and H) Expression of VE-cadherin in lung tissues and PMVECs isolated from CLEC5A^−/− mice and WT littermates after CLP challenge (biological replicates, n = 6 per group). (I) The permeability of the PMVEC monolayer was measured using FITC-labeled dextran by a transwell assay under LPS (10 μg/ml, 24 hours) (biological replicates, n = 4 per group). (J) TEER was assessed in LPS-challenged PMVECs (10 μg/ml, 4 hours) (biological replicates, n = 4 per group). (K) Quantification and representative images of immunofluorescence staining for VE-cadherin in PMVECs. Scale bar, 50 μm. (L) Expression of VE-cadherin in PMVECs with lentivirus-mediated CLEC5A overexpression (CLEC5A^oe) or knockdown (CLEC5A^sh) under LPS treatment (10 μg/ml, 24 hours) (biological replicates, n = 4 per group). The statistical significance between two groups was determined by an unpaired two-tailed t test and comparisons among more than two groups by one-way ANOVA analysis. Having observed the survival advantages of CLEC5A genetic deletion or endothelial-specific knockdown in CLP mice, we further assessed the bacterial loads in this model. At 24 hours after CLP, CLEC5A^−/− mice had an increase in the number of colony-forming units (CFUs) in the blood but decreases in the BALF and lungs, compared to WT littermates (fig. S7A). There was no significant difference in the blood, BALF, and lung bacterial contents from mice deficient in endothelial CLEC5A, compared to the controls (fig. S7B). These results indicate that the favorable effect of endothelial CLEC5A deficiency on the survival is likely not associated with bacterial clearance or spreading in the setting of CLP-induced abdominal sepsis. The mechanism underlying the therapeutic potential of targeting endothelial CLEC5A in sepsis needs to be unveiled and is the focus of this study. CLEC5A deficiency ameliorates pulmonary endothelial barrier dysfunction during inflammatory stimulation Pulmonary microvascular albumin leakage was determined by using Evans blue (EB) dye after CLP or sham operation ([125]Fig. 4, E and F). CLP induced microvascular leakage of EB-labeled albumin in the lungs of both CLEC5A^−/− transgenic mice and WT littermates. Mice lacking CLEC5A had less pulmonary albumin leakage than WT littermates after CLP challenge. CLP induces endothelial barrier dysfunction and resulting hyperpermeability and vascular leakage, which leads to albumin leakage and inflammatory cytokine and cell infiltration into the endothelium, thus causing tissue damage ([126]Fig. 3). The expression of vascular endothelial (VE)–cadherin, an endothelial cell junctional protein, was examined to assess endothelial integrity. In line with the in vivo data, the expression of VE-cadherin in lung tissues and primary PMVECs was increased in CLEC5A^−/− mice ([127]Fig. 4, G and H). We next investigated the dysfunctional cellular mechanisms underlying the therapeutic potential of CLEC5A in sepsis-associated lung injury. Under pathological conditions, dysfunction of vascular endothelial cells causes barrier failure ([128]10). Whether CLEC5A contributes to PMVEC dysfunction was explored using inflammatory stimulation with LPS. PMVECs were isolated from WT littermates and identified by CD31-positive immunofluorescence staining (fig. S2, C and D). The expression of CLEC5A was significantly increased in PMVECs upon LPS (10 μg/ml) stimulation (fig. S2, E and F). The barrier integrity and function of the PMVEC monolayer were determined during LPS-induced inflammation, and vehicle treatment served as the negative control. LPS stimulation impaired PMVEC barrier function, as shown by increased cell permeability of 1.75 fold in WT littermates ([129]Fig. 4I). Deletion of CLEC5A protected PMVECs against LPS-induced hyperpermeability. The trans-endothelial electrical resistance (TEER) of PMVECs was lowered in the presence of LPS, and CLEC5A-deficient PMVECs were resistant to LPS-induced TEER loss ([130]Fig. 4J). PMVECs challenged by LPS expressed less immunofluorescence for VE-cadherin, and genetic deletion of CLEC5A restored VE-cadherin level to a certain extent ([131]Fig. 4K). Lentivirus-mediated gene delivery and knockdown was applied in vitro with RNA interference (RNAi) technology, and the efficiency of targeting CLEC5A was validated in PMVECs (fig. S6C). In WT PMVECs, VE-cadherin expression was reduced by overexpression of CLEC5A and increased by knockdown of CLEC5A under LPS stimulation ([132]Fig. 4L). In consistent with the in vivo data, deletion of CLEC5A had no effect on the barrier integrity and function of PMVECs under sham or vehicle treatment ([133]Figs. 3 and [134]4). CLEC5A was lowly expressed in normal lung tissues and PMVECs ([135]Fig. 1, D to F). In comparison, CLEC5A was highly expressed during CLP or LPS-induced inflammation and closely associated with endothelial barrier dysfunction and lung injury. We suggest a pathological role of endothelial CLEC5A in sepsis-associated lung injury. CLEC5A promotes endothelial adhesion and transmigration of monocytes From the scRNA-seq analysis of mouse lungs, we observed that the number of monocytes was significantly lower in the septic lungs of CLEC5A^−/− mice compared to those of WT littermates ([136]Fig. 5A). UMAP analysis visualized the difference in monocyte clusters from CLEC5A^−/− and WT lungs. A total of 5786 single cells were identified as monocytes based on clustering analysis and cell type characterization, 3581 of which were obtained from WT littermates (n = 3) and 2205 of which were from CLEC5A^−/− mice (n = 3). Consistently, the proportion of monocytes in total immune cells was significantly decreased in CLEC5A-deficient lungs. Pulmonary infiltration of inflammatory cells, including monocytes, was increased in the setting of CLP-induced lung injury, which was ameliorated by genetic deletion of CLEC5A in mice ([137]Fig. 2B). Moreover, the expression levels of adhesion molecules were generally down-regulated in CLEC5A-deficient lung-derived endothelial cells, specifically in the capillary subtype of endothelial cells ([138]Fig. 5B). Among these adhesion molecules, down-regulation of intercellular adhesion molecule 1 (ICAM-1) and vascular cell adhesion molecule 1 (VCAM-1) was the most significantly affected by deletion of CLEC5A. According to gene set enrichment analysis (GSEA), we found multiple GO–Biological Process (BP) that are associated with cell adhesion were significantly enriched in the endothelial cells ([139]Fig. 5C), such as the regulation of cell-matrix adhesion (P = 0.0036) and the negative regulation of leukocyte adhesion to vascular endothelial cell (P = 0.0021). These cell adhesion–related GO pathways were mainly enriched by the DEGs that were up-regulated in CLEC5A^−/− endothelial cells. These findings suggest that endothelial cells may contribute to prevention of monocyte influx into the lung, thus protecting the lung from septic infection. Fig. 5. CLEC5A promotes endothelial adhesion and transmigration of monocytes in vitro. [140]Fig. 5. [141]Open in a new tab (A) UMAP plot of monocyte clusters in WT lungs (n = 3) and CLEC5A^−/− lungs (n = 3) and percentage in immune cells. (B) Dot plot showing the expression of adhesion molecules in each subtype of endothelial cells. (C) GSEA of the DEGs in the endothelial cells between CLEC5A^−/− and WT lungs showing significant enrichment of cell-adhesion related-GO pathways, in which the up-regulated DEGs were mainly enriched. The significance was determined by P value with false discovery rate (FDR) < 0.25. (D) Mouse PMVECs were isolated and challenged by LPS (10 μg/ml) for 24 hours. (E) Levels of MCP-1 and VCAM-1 in the supernatant by ELISA. (F) Relative mRNA expression of MCP-1 and VCAM-1 in PMVECs. (G) Overview of in vitro adhesion and trans-endothelial migration assays in mouse PMVECs and HUVECs. (H) Mouse PBMCs were labeled by BCECF-AM, and the adherent monocytes to PMVECs were observed under a fluorescence microscopy. Scale bar, 50 μm. Monocyte-endothelial adhesion was quantified as the percentage of fluorescence intensity relative to the WT group. (I) Transmigration of mouse PBMCs across the PMVEC monolayer was shown as the percentage of transmigrated cells relative to the WT group by MTT assay. Lentivirus-mediated gene delivery for CLEC5A (CLEC5A^oe) or shRNA targeting CLEC5A (CLEC5A^sh) were carried out in HUVECs. (J) Adhesion of THP-1 to HUVECs was shown as fluorescence and quantified as the percentage of fluorescence intensity relative to the control group. Scale bar, 50 μm. (K) Migration of THP-1 through the HUVEC monolayer shown as the percentage of transmigrated cells relative to the control group. The statistical significance was determined by an unpaired two-tailed t test (or with Welch’s correction) for comparison between two groups and by one-way ANOVA test for comparisons among four groups (biological replicates, n = 4 per group for each experiment). We here explored the role of CLEC5A in leukocyte-endothelial interactions and assessed the interaction of vascular endothelial cells and monocytes in vitro. PMVECs isolated from CLEC5A^−/− mice and WT littermates were stimulated by LPS for modeling sepsis-induced immune/inflammatory dysfunction in the endothelium ([142]Fig. 5D). The extravasation of immune cells is partially attributed to functional alterations in vascular endothelial cells, such as phenotypic changes toward pro-adhesion and production of adhesion molecules. The cellular mRNA expression and protein secretion of PMVEC-derived chemokine MCP-1 and adhesion molecule VCAM-1 were increased with LPS stimulation ([143]Fig. 5, E and F). The production of these LPS-induced molecules was lowered in PMVECs lacking CLEC5A. Loss of CLEC5A had no effect on the expression and secretion of MCP-1 and VCAM-1 by PMVECs in the absence of LPS. After exposure to LPS (10 μg/ml) for 5 hours, PMVECs were cocultured with fluorescence probe (BCECF-AM)–labeled mouse PBMCs for 1 hour, and decreased adherent monocytes were observed in PMVECs lacking CLEC5A with fluorescence microscopy ([144]Fig. 5, G and H). Deletion of CLEC5A caused more than a 40% reduction in monocyte-endothelial adhesion after LPS stimulation. Similarly, the transmigration of PBMCs across LPS-challenged PMVECs had a 50% decrease in the absence of CLEC5A ([145]Fig. 5I). We further performed gain-of-function and loss-of-function experiments for CLEC5A in PMVECs with the use of lentiviral vectors (fig. S8A). Consistent with the above results, knockdown of CLEC5A in PMVECs decreased the production of MCP-1 and VCAM-1 (fig. S8, B and C) and inhibited monocyte adhesion (fig. S8D) and trans-endothelial migration (fig. S8E). Conversely, overexpression of CLEC5A promoted LPS-induced transmigration of PBMCs across PMVECs. Next, the regulation of CLEC5A on human monocyte-endothelial cell interactions was investigated in vitro. The mRNA and protein expression of CLEC5A was increased in both human umbilical cord endothelial cells (HUVECs) and human pulmonary microvascular endothelial cells (HMVECs) under LPS stimulation (fig. S2, G and H). Knockdown of CLEC5A in HUVECs suppressed LPS-induced endothelial adhesion and transmigration of human monocytes THP-1 ([146]Fig. 5, J to K, and fig. S6D). In contrast, forced overexpression of CLEC5A increased THP-1 adhesion to HUVECs and trans-endothelial migration. CLEC5A contributes neutrophil-endothelial interaction Damage to endothelial cells and resulting barrier dysfunction during sepsis stimulate more immune cells, such as neutrophils, to cross the endothelium, leading to local inflammation and tissue damage ([147]9). These in vivo experiments show that genetic deletion of CLEC5A inhibited the extravasation of neutrophils and influx into the lungs in mice with CLP-induced polymicrobial sepsis. LPS stimulation induced expression and release of CXCL5 and ICAM-1 that contribute to neutrophil recruitment and adhesion ([148]Fig. 6, A to D). Knockdown of CLEC5A significantly lowered the production of CXCL5 and ICAM-1 in LPS-challenged PMVECs, whereas it had no effect in the absence of LPS. Down-regulation of these factors expressed by CLEC5A-deficient PMVECs might account for reduced transmigration of neutrophils. By comparing PMVECs isolated from CLEC5A^−/− transgenic mice and WT littermates, we found CLEC5A deficiency caused a 40% reduction in the transmigration of neutrophils across the PMVEC monolayer under LPS-stimulated inflammation ([149]Fig. 6, E and F). The involvement of CLEC5A in neutrophil-endothelial cell interaction was further validated by gene knockdown and overexpression experiments using lentiviral transduction. PMVECs with lentivirus-mediated CLEC5A knockdown induced decreases in CXCL5 and ICAM-1 expression and neutrophil transmigration under LPS stimulation ([150]Fig. 6, B, D, and G). In contrast, viral delivery of transgene CLEC5A increased trans-endothelial migration of neutrophils and PMVEC-derived CXCL5 and ICAM-1 levels. In the absence of LPS, loss of CLEC5A did not induce damage to endothelial cells either in terms of permeability or barrier function ([151]Figs. 4 to [152]6). Fig. 6. CLEC5A promotes trans-endothelial migration of neutrophils in vitro. [153]Fig. 6. [154]Open in a new tab Mouse PMVECs were exposed to LPS (10 μg/ml) for 24 hours, and the production of CXCL5 and ICAM-1 was detected in the presence or absence of CLEC5A. Expression of CXCL5 and ICAM-1 by real-time PCR in (A) PMVECs from CLEC5A^−/− transgenic mice and (B) in PMVECs with lentivirus-mediated overexpression (CLEC5A^oe) or knockdown of CLEC5A (CLEC5A^sh). (C and D) Respective levels in the supernatant by ELISA. (E) Migration of neutrophils across LPS-challenged PMVECs was determined by a transmigration assay. Peripheral blood granulocytes were obtained from normal C57BL/6J mice, and granulocytes that transmigrated across the PMVEC monolayer were identified as neutrophils under phase contrast microscopy. Representative images of transmigrated neutrophils to cross (F) PMVECs from CLEC5A^−/− mice and (G) PMVECs infected with lentivirus-expressing CLEC5A (CLEC5A^oe) or shRNA targeting CLEC5A (CLEC5A^sh). Scale bars, 50 μm. Trans-endothelial migration of neutrophils was quantified as the percentage of transmigrated cell number relative to WT or control group. Cells with LPS stimulation but without infection were served as the controls. The statistical significance was determined by an unpaired two-tailed t test for comparison between two groups and by one-way ANOVA test for comparisons among four groups (biological replicates, n = 4 per group for each experiment). CLEC5A is critical in bacterial pneumonia induced by E. coli intratracheal instillation To further determine whether the role of CLEC5A in lung inflammation was restricted to abdominal sepsis, we established a mouse model of bacterial pneumonia. CLEC5A^−/− mice and WT littermates received an intratracheal instillation of Gram-negative bacterium E. coli. Similar to the results in the CLP model, E. coli–induced pulmonary edema and pathological lesions in lung tissues by H&E staining were alleviated in CLEC5A^−/− mice ([155]Fig. 7, A and B) and endothelial CLEC5A-deficient mice ([156]Fig. 7, C and D). Global or endothelial-specific deletion of CLEC5A increased the expression of VE-cadherin in lung tissues after E. coli infection ([157]Fig. 7, E and F). Consistently, pulmonary microvascular albumin leakage ([158]Fig. 7, G and H) as well as inflammatory cell infiltration ([159]Fig. 7, I and J) and cytokine levels in BALF ([160]Fig. 7, K and L) were reduced in the absence of endothelial CLEC5A. Either global deletion or endothelial-specific knockdown of CLEC5A had advantages in host defense against E. coli infection and the resulting lung inflammation and vascular injury. Furthermore, the bacterial loads were determined in E. coli–induced pneumonia model. The bacterial CFUs were increased in the blood but decreased in the BALF and lungs in CLEC5A^−/− mice after E. coli intratracheal inoculation (fig. S7C), whereas there was no significant difference in mice with endothelial CLEC5A knockdown (fig. S7D). Fig. 7. Endothelial CLEC5A deficiency protects mice against E. coli–induced pneumonia. [161]Fig. 7. [162]Open in a new tab Mice were intratracheally inoculated with 1.5 × 10^7 CFUs of E. coli and euthanized after 16 hours. The lungs were collected from CLEC5A^−/− mice and endothelial-deficient mice (CLEC5A^Tie1-sh). (A and C) Ratio of lung wet-to-dry weight. (B and D) Representative images of H&E staining of lung tissues and quantitative analysis of lung injury severity. Scale bars, 100 μm. (E and F) Relative expression of VE-cadherin in lung tissues. (G and H) Mice received EB at 16 hours after bacterial inoculation, and pulmonary microvascular albumin leakage (μg EB/g lung per minute) was determined after 30 min. (I and J) BALF cell counts of leukocyte, neutrophil, and monocyte. (K and L) BALF levels of TNF-α and IL-6. The statistical significance between two groups was determined by an unpaired two-tailed t test, and the Mann-Whitney test was used for lung injury score (biological replicates, n = 6 per group for each experiment). Exploration of the downstream targets of CLEC5A responsible for endothelial dysfunction in inflammatory lung injury Having observed the role of CLEC5A in the protection of endothelium against barrier failure during inflammation as well as the contribution of endothelial CLEC5A to sepsis-induced mortality, we further explored the downstream targets of CLEC5A to uncover potential molecular mechanisms in the endothelium. An in-depth scRNA-seq transcriptomic analysis was then carried out in endothelial cells. Heatmaps reveal the total or specific DEGs in vascular and capillary endothelial cells from CLEC5A^−/− and WT lungs challenged by CLP ([163]Fig. 8A). Capillary endothelial cells were found to have the largest number of DEGs between CLEC5A^−/− and WT lungs ([164]Fig. 8B), implying that capillary endothelial cells play an important role in sepsis-associated lung injury. Analysis of scRNA-seq showed many overlapping DEGs in the vascular subtype of endothelial cells, and some specifically dysregulated in the capillary type of endothelial cells. These single-cell transcriptomic data demonstrate cell heterogeneity of the lung endothelium and endothelial cell hierarchy in CLEC5A-deficient mice following CLP challenge. Specifically, capillary endothelial cells isolated from septic lungs of CLEC5A^−/− mice display overlapping and unique features in aspects of endothelial dysfunction, sepsis pathophysiology, and the development of sepsis-associated lung injury. Fig. 8. Transcriptome changes in vascular/capillary endothelial cells by scRNA-seq analysis. [165]Fig. 8. [166]Open in a new tab An in-depth analysis of the DEGs was carried out in specific endothelial subtypes. (A) Heatmaps of the DEGs among five subtypes of endothelial cells between WT and CLEC5A^−/− lungs. The column is hierarchically ordered by the subtypes of endothelial cells (aerocyte, arterial, gCap, lymphatic, and venous) and then groups of WT and CLEC5A^−/−, with single-cell gene expression represented by each column. The vascular endothelial cells include subtypes of aerocyte, arterial, gCap and venous. The capillary endothelial cells include aerocyte and gCap of endothelial cells. (B) Number of the DEGs in capillary (gCap and aerocyte), arterial, venous, and lymphatic types of endothelial cells. (C) Violin plots showing the expression of SCARB1, PODXL, RAMP2, and FABP4 in capillary endothelial cells. The expression was further validated in primary PMVECs isolated from CLEC5A^−/− mice or WT littermates 12 hours post-CLP. (D) Representative blots showing SCARB1, PODXL, FABP4, and RAMP2 expression. (E and F) The relative protein and mRNA expression levels were analyzed (biological replicates, n = 4 per group). The statistical significance was determined by an unpaired two-tailed t test for comparison between two groups. We observed that some DEGs were significantly and specifically dysregulated in the vascular or capillary subtype of endothelial cells ([167]Fig. 8, A and B). Among them, we noticed a PRR, scavenger receptor class B membrane 1 (SCARB1), and an anti-adhesive factor, podocalyxin like (PODXL), the expression of which was significantly up-regulated ([168]Fig. 8C). Of note, fatty acid–binding protein 4 (FABP4) and receptor activity modifying protein 2 (RAMP2) were found to be significantly down-regulated in capillary endothelial cells from CLEC5A-deficient lung ([169]Fig. 8C). These potential targets downstream of CLEC5A were further confirmed in CLEC5A-deficient PMVECs. We found that the expression of SCARB1, PODXL, FABP4, and RAMP2 was not affected in PMVECs from CLEC5A^−/− mice under physiological conditions (fig. S9, A and B), whereas they were specifically targeted by CLEC5A upon CLP challenge. Deletion of CLEC5A increased endothelial expression of SCARB1 and PODXL and decreased RAMP2 and FABP4 at both protein and mRNA levels in mice subjected to CLP ([170]Fig. 8, D to F). Next, we examined whether these DEGs are involved in the therapeutic effect of CLEC5A deficiency in CLP-induced sepsis and lung injury. Endothelial specific knockdown of SCARB1 or PODXL and re-enforced expression of FABP4 or RAMP2 was carried out under the endothelial-specific promoter Tie1 by AAV9 in CLEC5A^−/− mice as previously described. Potential molecular mechanisms underlying endothelial CLEC5A-mediated lung inflammation and vascular injury in sepsis SCARB1 as a PRR is able to sense PAMPs and DAMPs, involves in host defense against microbial infection ([171]19), and has been reported to protect against polymicrobial sepsis and endotoxemia in mice ([172]20, [173]21). Inhibition of endothelial SCARB1 completely abolished the improvement in the survival rate observed in CLP-challenged CLEC5A^−/− mice ([174]Fig. 9A). All mice that lacked CLEC5A and endothelial SCARB1 died within 72 hours after CLP. Loss of endothelial SCARB1 aggravated CLP-induced lung injury ([175]Fig. 9E), along with increases in inflammatory cell (leukocyte, neutrophil, and monocyte) infiltration and inflammatory cytokine (TNF-α and IL-6) and chemokine (MCP-1) production in the BALF and lungs of CLEC5A^−/− mice ([176]Fig. 9, J, L, and N). These results suggest a key contribution of SCARB1 to CLEC5A-mediated barrier function in the endothelium and lung protection during septic infection. Fig. 9. Potential downstream targets of CLEC5A in CLP-induced lung injury. [177]Fig. 9. [178]Open in a new tab To validate SCARB1, PODXL, RAMP2, and FABP4 as potential targets of CLEC5A in vivo, endothelial-specific gene knockdown/overexpression was carried out under Tie1 by AAV9 through tail vein injection. (A to D) Survival rate of CLEC5A^−/− mice with endothelial knockdown/overexpression of target gene after CLP (biological replicates, n = 10 per group). The statistical significance between survival curves was determined by P value using the log-rank (Mantel-Cox) test. (E to G) H&E staining showing histological changes in lung tissues and the corresponding lung injury score (biological replicates, n = 6 per group). Scale bars, 100 μm. (H and I) Pulmonary microvascular albumin leakage represented as μg EB/g lung per minute (biological replicates, n = 6 per group). (J and K) Levels of TNF-α, IL-6, and MCP-1 in the BALF (biological replicates, n = 6 per group). (L and M) Levels of TNF-α, IL-6, and MCP-1 in the lungs (biological replicates, n = 6 per group). (N and O) Quantification of inflammatory cells in the BALF, including leukocyte, neutrophil, and monocyte (biological replicates, n = 6 per group). (P) HUVECs were coinfected with LV expressing CLEC5A^sh and PODXL^sh. The monocyte-endothelial adhesion and trans-endothelial migration assays were carried out under LPS stimulation. (Q) Adhesion of THP-1 to HUVECs (biological replicates, n = 4 per group). (R) Transmigration of THP-1 across HUVECs (biological replicates, n = 4 per group). The statistical significance was determined by an unpaired two-tailed t test (or with Welch’s correction) for comparison between two groups. FABP4 is a cytosolic lipid binding protein that plays a role in regulating the function of capillary endothelium and immune/inflammatory responses ([179]22, [180]23), and its inhibitor had therapeutic effects in experimental sepsis-induced ALI ([181]24). Genetic overexpression of endothelial FABP4 not only abolished the therapeutic effect of CLEC5A depletion but also aggravated CLP-induced mortality and lung injury in mice, as evidenced by 100% mortality within 60 hours after CLP ([182]Fig. 9B) and higher lung injury score at 12 hours after CLP ([183]Fig. 9F). Up-regulation of FABP4 increased inflammatory cytokine levels and infiltrated cells in CLP-challenged lungs of CLEC5A^−/− mice ([184]Fig. 9, K, M, and O). Investigation of FABP4 as a downstream effector of CLEC5A in the endothelium or a potential regulator in CLEC5A-mediated inflammatory signaling pathways is an area that requires further study. PODXL is widely expressed in vascular endothelium and is required for functional barrier formation and maintenance under resting and inflammatory conditions ([185]25). Following CLP, we found that deletion of endothelial PODXL increased CLP-induced mortality to a certain extent ([186]Fig. 9C) and significantly increased lung injury and vascular leakage in CLEC5A^−/− mice ([187]Fig. 9, G and H). Knockdown of PODXL enhanced leukocyte-endothelial adhesion and trans-endothelial migration of THP-1 in CLEC5A-deficient HUVECs ([188]Fig. 9, P to R). Endothelial specific knockdown of PODXL worsened lung injury in CLEC5A^−/− mice by inducing endothelial cell dysfunction and its pro-adhesive phenotype. Adrenomedullin is a potent vasodilator and contributes to anti-inflammation and antimicrobial infection in sepsis, although its vasodilatory effect could be detrimental ([189]26). RAMP2 acts as adrenomedullin-receptor accessory protein, and adrenomedullin-RAMP2 system is crucial in maintaining vascular integrity, endothelial barrier function and organ homeostasis ([190]27). Although there was a down-regulation of RAMP2 in the absence of CLEC5A, no effect on survival after CLP or pulmonary vascular leakage was observed in CLEC5A^−/− mice with overexpression of endothelial RAMP2 ([191]Fig. 9, D and I). Therefore, CLEC5A might not function via endothelial RAMP2 in the setting of CLP-induced lung injury and barrier dysfunction. DISCUSSION The immune profile of sepsis includes dysregulated immune responses to infection, resulting in a systemic inflammatory response ([192]28). In the initial phase of sepsis, the innate immunity is activated and initiates inflammation. Persistence of the inflammatory storm causes tissue damage, organ dysfunction, and subsequent death. Previous studies have shown that abdominal infection-induced sepsis in vivo, including intraperitoneal injection of LPS-induced endotoxemia and CLP-induced polymicrobial sepsis, results in severe peritoneal inflammation and multiple organ injury ([193]29, [194]30). The hyperinflammatory response is closely correlated with early death following CLP or LPS. Abdominal sepsis often results in lung injury and ARDS and ultimately death ([195]31). It has been well known that C-type lectins function in host defense against pathogens and immune homeostasis. CLEC5A, a C-type lectin, is identified as a PRR for several types of viruses and bacteria. Studies have shown that glycan-lectin interactions activate immune cells and trigger inflammatory responses ([196]32). Bacterial LPSs directly interact with cells via PRRs, including C-type lectins, and mediate intracellular signalosome formation and activation. C-type lectins, as glycan-binding receptors, recognize glycan structures on viruses, bacteria, and other microorganisms and initiate immune and inflammatory responses ([197]33, [198]34). It has been reported that CLEC5A interacts with the LPS of P. aeruginosa and mediates pneumonia in mice ([199]14). Blockade of CLEC5A or CLEC5A-mediated signaling inhibited pro-inflammatory responses and increased survival of mice exposed to P. aeruginosa. Deletion of CLEC5A in neutrophils inhibited pro-inflammatory cytokine production, but mice lacking CLEC5A had a higher susceptibility to Listeria infection ([200]13), which might be associated with failed homeostatic immunity. These observations suggest that CLEC5A acts as an inflammatory signaling receptor/sensor, and genetic deletion of CLEC5A interferes with sensing the danger signals or the signal transduction of LPS or CLP responsiveness. In our study, deletion of CLEC5A greatly improved survival of mice challenged by LPS or CLP, as well as prevented the development of lung inflammation and injury. These benefits might be attributed to CLEC5A’s role in modulating dysregulated immune and inflammatory responses. CLEC5A-deficient mice had increased number of bacterial CFUs in the blood but decreased number in the BALF and lungs. CLEC5A is critical for neutrophil extracellular trap formation and controlling bacterial spreading ([201]13, [202]14). The innate immune response to bacterial infection could be impaired in the CLEC5A^−/− mice and thus the failure of bacterial clearance, resulting in higher bacterial contents in the blood. Decreased bacterial loads in the BALF and lungs might be associated with less recruitment and infiltration of immune/inflammatory cells to the lungs, whereas this could be attributed to reduced production and secretion of inflammatory cytokines and chemokines in the lung. The inability of CLEC5A^−/− immune cells in controlling the spread of bacteria in the lungs might facilitate the spreading to the blood. These observations suggest the importance of CLEC5A in host defense against microbial infection, whereas the survival benefits of CLEC5A deficiency in sepsis models maybe not due to control of bacterial infection and spreading. Further exploration of the underlying mechanism and the relevance to mortality in sepsis is needed. There is currently a lack of evidence to support the immunopathologic mechanism of sepsis. A hypothesis proposes that barrier failure contributes to lethal organ dysfunction in sepsis ([203]2). Homeostatic organ function is dependent on barriers between tissue compartments. These barriers, maintained by endothelial and epithelial cells, support organ function and are responsible for regulating the movement of molecules across tissues. Impaired homeostatic regulation of barrier function is evident in clinical and experimental sepsis and associated with the development of lung injury ([204]35). Abnormalities in the lungs following CLP, including edema formation, vascular leakage, inflammatory cell infiltration, and pro-inflammatory cytokine overproduction, are likely attributed to a failure of the alveolar-capillary barrier, which is responsible for triggering ARDS, and its subsequent high mortality. Along with failed pulmonary homeostasis and lethal lung injury is an up-regulation of CLEC5A in the lungs. Current research focuses mostly on CLEC5A produced by myeloid cells and has shown that knockdown of CLEC5A inhibits the production of pro-inflammatory cytokines and attenuates lung injury under infectious conditions ([205]14, [206]36). Consistently, we showed that mice that lack CLEC5A were resistant to lung injury and inflammation, and thus mortality in the setting of CLP-induced polymicrobial sepsis. We further extend the study of CLEC5A to pneumonia and inflammatory lung injury. In the E. coli–induced pneumonia model, there were less lung inflammation and vascular leakage in CLEC5A^−/− mice, indicating deletion of CLEC5A protected the lung against E. coli infection. These observations were similar to that in the CLP model. Here, we asked whether and how CLEC5A affects cellular and organ dysfunction in the progression of microbial infection to lung injury and mortality. The endothelium has been well characterized in the control of local infection through releasing inflammatory mediators and recruiting leukocytes. During sepsis, the endothelium is activated by exogenous pathogen-associated molecular patterns (PAMPs) and endogenous damage-associated molecular patterns (DAMPs) via PRRs ([207]37). Its activation is beneficial for the local immune response to localized infection but detrimental under systemic activation in sepsis ([208]38). Dysfunction of the endothelium is responsible for lung injury, and targeting the endothelium is a promising approach to limit sepsis-associated organ failure and, consequently, death ([209]39). Selective delivery of therapeutic molecules can be achieved using liposomes that are designed to encapsulate pharmacologically active entities. Recent studies have shown endothelial cell specific delivery by liposomes modified with antibodies against VCAM-1 ([210]40, [211]41). Dysregulation of CLEC5A was found in pulmonary endothelial cells during sepsis and inflammation. CLEC5A deficiency prevented the triggering of lung injury in response to microbial infection and decreased mortality after CLP, and notably, specific deletion of endothelial CLEC5A had a similar outcome. The contribution of endothelial CLEC5A to poor outcomes in mice with polymicrobial sepsis suggests the pathogenic role of CLEC5A in the endothelium and the lungs in the defense against septic infection. CLEC5A^−/− transgenic mice seemed to have a better survival than endothelial-specific knockdown in mice following CLP. The role of CLEC5A, as a PRR, has been well described in myeloid cells and involvement in immune and inflammatory responses ([212]42). Previous studies have suggested the importance of CLEC5A in innate immunity against bacterial infection in vivo ([213]13, [214]14). The effect of global knockout (CLEC5A^−/−) in sepsis-induced lung injury and mortality might be due to CLEC5A from myeloid cells. In both CLP-induced abdominal sepsis and E. coli–induced pneumonia models, genetic deletion of CLEC5A might disturb innate immunity in response to bacterial infection and thus the ability of bacterial clearance and the control of bacterial spreading. In contrast, endothelial CLEC5A deficiency in mice had no effect on bacterial loads in the blood, BALF, and lung but inhibited the development of sepsis-associated lung injury and E. coli pneumonia. The lung protection of CLEC5A deletion may be due to inhibition of inflammatory storm in the lung and maintenance of endothelial barrier function. Our study identifies an immunological mechanism of CLEC5A in the endothelium by which CLEC5A mediates local inflammatory responses to microbial infection and contributes to the progression of lung injury. CLEC5A-deficient endothelium was resistant to sepsis-induced barrier dysfunction and thus inhibited immune cell influx into the lungs. Endothelial CLEC5A might be a target for therapeutic approaches to sepsis and its associated lung injury. Targeting endothelial CLEC5A with the use of Tie1 promoter could be a limitation of the study because Tie 1 is also expressed by immature hematopoietic cells. Endothelial-specific interference of CLEC5A expression/function is an issue need to be solved in the further research, as well as the specificity of AAV9-mediated deficiency or overexpression approach to the lungs. An alternative to selectively target CLEC5A in endothelial cells would provide a better understanding of CLEC5A’s role in the endothelium. The cellular function and mechanisms of CLEC5A in the setting of bacterial infection and the resulting inflammatory lung injury, and the possibility of other tissue or cell types that may be involved, deserve an in-depth investigation. Sepsis-induced damage to endothelial cells leads to impaired structure and function of the endothelium. PMVEC permeability impairment underlies pulmonary microvascular leakage. This leakage induces extravasation of albumin into the lungs, where it accumulates and produces edema ([215]10). Consistent with in vivo findings, deletion of CLEC5A protected PMVECs against LPS-induced barrier failure by maintaining endothelial integrity and hyperselective function. This endothelial protection accounted for decreased exudation and vascular leakage in CLEC5A-deficent mice following CLP and prevented lung injury. Leakage across the alveolar-capillary barrier allows the passage of pro-inflammatory cytokines (such as TNF-α and IL-6) that are released into the bloodstream from other inflammation sites to the pulmonary endothelium. It has been established by Tracey et al. that TNF produced by activated myeloid cells, mainly neutrophils and macrophages, is crucial in this immunopathology and sufficient to mediate lethal tissue injury ([216]43). TNF-α induces capillary leakage and abnormal vasodilation, and contributes to vascular changes and the progression of sepsis ([217]42). IL-6 released by leukocytes impairs intercellular junctions by decreasing VE-cadherin, leading to endothelial barrier disruption ([218]44). Loss of CLEC5A maintained cohesion and organization of PMVECs and inhibited LPS-induced intercellular gaps by increasing the endothelial junction protein VE-cadherin. These inflammatory mediators cause damage to endothelial cells, such as inducing cell apoptosis ([219]45). Apoptosis of endothelial cells leads to barrier failure and interstitial leakage, and we found that CD31-positive endothelial cells were decreased in the lungs of mice subjected to CLP. These adverse consequences in the endothelium were diminished by deletion of CLEC5A, and its contribution to maintaining barrier function could be due to the improvement of endothelial cell survival. On the other hand, apoptotic cells favor the expression of cytokines, chemokines and adhesion molecules and thus recruitment of leukocytes, which, in turn, amplifies local immune responses and propagates endothelial dysfunction ([220]46). Along this line, CLEC5A deficiency prevented against pulmonary inflammation and cell infiltration in the setting of CLP-induced endothelial dysfunction. These findings suggest that the lung protection observed from CLEC5A deficiency may be achieved through preventing CLP-induced endothelium failure, thereby improving survival. Under severe conditions, defects of the endothelial barrier are exacerbated and leaks are enlarged to enable infiltration of blood immune cells, including leukocytes, monocytes, and neutrophils, into the tissues. Neutrophils account for 50 to 70% of white blood cells in humans and act as the first line against infection in the innate immune system ([221]47). Excessive accumulation of neutrophils and recruitment of leukocytes to the tissues exacerbates local inflammation and vascular leakage ([222]8). CLEC5A-deficient PMVECs and HUVECs reduced trans-endothelial migration of monocytes and neutrophils across the endothelium during inflammation. Mice that lacked CLEC5A consistently had less pulmonary infiltration of immune cells and decreased inflammation following CLP. Regulation of immune cell recruitment and subsequent transmigration across the endothelium is important in the control of local inflammation and tissue injury during sepsis ([223]48). We demonstrate that CLEC5A is crucial for barrier function of pulmonary endothelium in cross-talk with immune cells, and targeting CLEC5A is promising as protection against immune cell influx and immune/inflammatory dysregulation in sepsis-induced lung injury. Sepsis results in a “gain-of-function” in endothelial cells, including pro-inflammatory and pro-adhesive properties. Upon engagement of PRRs, endothelial intracellular signaling is activated, followed by consecutive activation of transcription factors such as nuclear factor κB subunit 1 (NFKB1/NF-κB) ([224]49). The transcription of downstream genes consequently increases, including pro-inflammatory cytokines/chemokines and cell adhesion molecules. Expression of these mediators regulates leukocyte trafficking to the surrounding tissues and interaction with endothelial cells before transmigration ([225]50). Under inflammatory stimulation, PMVECs release MCP-1 and VCAM-1, which favor monocyte recruitment and adherence, and produce CXCL5 and ICAM-1 that contribute to neutrophil influx into the lungs. CLEC5A-deficient endothelial cells negatively regulated adhesion of monocytes and neutrophils by inhibiting gene transcription and secretion of these mediators. Our findings support the notion that CLEC5A mediates endothelial activation and intracellular cytokine production, resulting in functional alterations in endothelial cells during sepsis. Improved understanding of the signaling pathways downstream of CLEC5A-mediated endothelial cell alterations would provide insights into therapeutic strategies targeting CLEC5A in sepsis. Together with these observations, CLEC5A is involved in endothelial phenotypic modification and pro-inflammatory responses upon septic infection. Sepsis-related endothelial change is a double-edged sword, and it is challenging to distinguish between its adaptive response and dysfunction. More in vivo and in vitro studies to understand the underlying mechanisms by which CLEC5A mediates endothelial activation and/or dysfunction in response to infection are still necessary, and differences in organ-specific endothelial cell responses need to be considered. A promising therapeutic strategy for sepsis-associated organ failure is the targeting of the pathways involved in endothelial dysfunction without affecting appropriate activation. Limiting vascular leakage protects the endothelium from inflammatory stimuli and is considered a therapeutic approach to experimental sepsis. However, these effects may also impair leukocyte recruitment and affect the local immune response, which is harmful in defense against pathogens ([226]51). Filewod and Lee and recent studies have reported that vascular leakage can be regulated without affecting immune responses and pathogen clearance ([227]52). In-depth exploration is needed to distinguish the pathways by which CLEC5A regulates leukocyte trafficking and vascular leakage/permeability. In the present study, we demonstrated that knockdown of endothelial CLEC5A increased survival of mice with polymicrobial sepsis or endotoxemia and alleviated lung injury and inflammation caused by bacterial infection. Deletion of CLEC5A protected against pulmonary endothelial cell dysfunction and barrier failure during inflammation, and inhibited leukocyte-endothelial adhesion and transmigration, thus maintaining pulmonary homeostasis. We suggest endothelial CLEC5A is a key protein responding to sepsis-causing bacterial infection, via mediating barrier function and endothelial-immune cell cross-talk. This work extends the potential of targeting CLEC5A in the treatment of inflammatory lung injury due to an infection such as bacterial pneumonia and sepsis and opens the field for mechanistic and therapeutic implications of CLEC5A in the progression of sepsis in humans exposed to a variety of infectious conditions. MATERIALS AND METHODS Ethics approval All animal experiments were approved by the Institutional Ethics of Animal Care and Use Committee at Shengjing Hospital of China Medical University (2021PS637K). Animals C57BL/6 WT mice were purchased from Charles River (Wilmington, Massachusetts, USA). By using the online random number generator ([228]https://iikx.com/tool/radom.html), mice were randomly assigned into groups with no blinding based on study design. A priori power analysis was performed to determine the group size using an online calculator PowerAndSampleSize ([229]http://powerandsamplesize.com/), based on a two-sample and two-sided equality to compare two means with a power (1-β, Type II error) of 0.8 and an α (Type I error) of 0.05. Equal numbers of male and female mice were included in each animal experiment. All mice were used and euthanized at the age of 8 to 12 weeks. All animal experiments were approved by the Institutional Ethics of Animal Care and Use Committee at Shengjing Hospital of China Medical University (2021PS637K). The CLEC5A gene (NCBI Gene ID: 23845) is located on mouse chromosome 6 (Chr: 40551832-40562739) with a full length of 10,908 base pairs (bp). CLEC5A genetic knockout mice were generated on a C57BL/6 background and obtained from Wanlei Life Sciences Co. Ltd. (Shenyang, China). Briefly, two single guide RNAs (sgRNAs) were designed for targeting the region from exon 2 to exon 5 of CLEC5A (NCBI Reference Sequence: [230]NM_021364, knockout region size: 8501 bp). The target sequences are 5′-TGGTTCCGTAGCTCTCCGTG-3′ (#1) and 5′-AATGAGCCTTTGATCTACGG-3′ (#2). A mixture of CRISPR-Cas9 recombinant protein and sgRNAs was coinjected into zygotes for generating CLEC5A^−/− mice. WT littermates were used as the control. Model of experimental sepsis Mice of a C57BL/6 background, 8 to 10 weeks old, were used for the modeling. Two mouse models of sepsis were established based on modified previous methods. For induction of polymicrobial sepsis, mice were subjected to CLP with two punctures using a 25-gauge needle as previously described ([231]53, [232]54). Sham mice underwent the same operation except for cecal ligation or puncture. For endotoxemia, mice received a single intraperitoneal injection of LPS (10 mg/kg; E. coli 055:B5; L2880#, Sigma-Aldrich, St. Louis, MO, USA). In vehicle group, mice received same amount of saline. For survival analysis, N = 10 mice were included in each group and the survival was recorded every 12 hours for 1 week. N = 6 mice per group were applied in all other experiments. Lung histological assessment and immunofluorescence staining The lungs were isolated, photographed, and weighted (wet and dry). Paraffin-embedded sections with a thickness of 5 μm were prepared and used for H&E staining. Severity of lung injury was blindly assessed based on inflammation, edema, hemorrhage, and alveolar septal thickening using a 0-4 point scoring system, individually ([233]55). Lung injury score was determined as the sum of each score from above four criteria. N = 6 mice per group were for histological analysis, including three randomly selected slides per mouse with three random fields for each slide. Immunofluorescence for CLEC5A (AF1639#, R&D system, Minneapolis, MN, USA) and CD31 (ab222783#, Abcam, Cambridge, UK) was performed according to a standard protocol with the use of 4′,6-diamidino-2-phenylindole (DAPI) for nuclear counterstaining. For studying the expression pattern of CLEC5A in specific endothelial cell subtypes, we conducted double immunofluorescent staining for CLEC5A and GJA5 (680144#, Zenbio, Chengdu, China), VWF (AF3000#, Affinity Biosciences, China), EDNRB (DF7104#, Affinity Biosciences), PTPRB (AP58204#, Abcepta, Jiangsu, China), or PROX1 (bs-2774R#, Bioss, Beijing, China) in lung tissues. BALF collection The left bronchus was ligated, and BALF was collected by flushing the lung with 0.4 ml of phosphate-buffered saline (PBS) (B548117#, Sangon Biotech, Shanghai, China) via the tracheal cannula, three times per mouse. Inflammatory cells in the BALF were collected, and total leukocyte counts were determined using a hemocytometer. Differential cell counts for neutrophils and monocytes were performed with Wright-Giemsa stain (D010#, Nanjing Jiancheng Bioengineering Institute, Nanjing, China) following morphological criteria. Microvascular albumin leak Pulmonary microvascular permeability was evaluated by intravenous injection of EB dye as previously described ([234]56). Briefly, mice received EB dye (50 μg/g) via tail vein 30 min prior to euthanasia. PBS-perfused lung was collected and the EB dye was extracted by formamide incubation. The EB content was measured, and the leakage of EB-labeled albumin was calculated as mg EB/g lung per minute. Model of pneumonia C57BL/6 mice were challenged by 1.5 × 10^7 CFUs of E. coli (American Type Culture Collection, strain 25922) in 50 μl of PBS through intratracheal instillation. The control mice received the same amount of PBS. Mice were euthanized 16 hours after bacterial infection. The blood, BALF, and lungs were collected, and N = 6 mice per group were included in each experiments. Determination of CFUs The bacterial loads were assessed in diluted samples of the peripheral blood, BALF, and lungs. In the CLP model, the samples were collected 24 hours after surgery. In the E. coli–induced pneumonia model, the samples were collected 16 hours after bacterial inoculation. For visceral CFU determination, the lungs were weighted and homogenized (0.1 g/ml in sterile PBS). After serial dilutions in sterile PBS (1:10, 1:100, or 1:1000), 100 μl of the diluent was plated on tryptoic soy agar plates (T8650#, Solarbio). Bacterial CFUs were counted after incubation at 37°C for 24 hours. The results for the blood and BALF bacterial counts are presented as CFUs/ml, and for the lung is CFUs per gram tissue. Single-cell RNA sequencing Lung tissues were obtained from CLEC5A^−/− and WT mice subjected to CLP. Single-cell preparation, scRNA-seq, and bioinformatics analysis were conducted by LC-Bio Technology (Hangzhou, China) ([235]http://lc-bio.com/tech-service-detail?cid=18&id=48). Briefly, the single-cell suspension was prepared from lung tissues according to the standard protocol. Cell viability (above 85%) was determined by Trypan blue exclusion, and samples were prepared at the concentration of 0.7 × 10^6~1.2 × 10^6 cells/ml. Single-cell suspensions were processed using Chromium Single Cell Solution Kit (10×Genomics) according to the manufacturer’s instructions. It was followed by cDNA amplification, and library construction and running on the NovaSeq 6000 system for Illumina sequencing with 2 × 10^4~4 × 10^4 mean reads per cell. The sequencing data were analyzed using the Cell Ranger software (V7.0.0) ([236]https://support.10xgenomics.com/single-cell-gene-expression/softw are/overview/welcome) and aligned to Ensembl Genome GRCm39 ([237]ftp://ftp.ensembl.org/pub/release-105/fasta/mus_musculus/dna). The output data were further analyzed by the Seurat software (V4.1.0) ([238]https://satijalab.org/seurat/). Briefly, a total of 64,581 single cells were captured from six mouse lungs following 10×Genomics Chromium Single Cell processing. The threshold for data quality control was set as follows: genes expressed per cell > 500 (gene number), mitochondrial genome content rate < 25%, and removal of multiplet cells by DoubleFinder. Overall, 55,177 single cells were obtained after filtration. For clustering and visualization, the data were normalized using the LogNormalize method by Seurat, followed by PCA. The top 20 PCs were used for subsequent dimensional reduction and clustering analysis. A UMAP analysis was carried out to visualize the identified cell clusters. The top marker genes that highly expressed in each cluster were identified using the FindAllMarkers function of Seurat. Within a default process, the marker gene was selected according to the following thresholds: genes expressed in >10% cells in a cluster, P value ≤ 0.01, and log[2]FC ≥ 0.26. Viral vector and gene transduction The loss-of-function study for CLEC5A was carried out by RNAi. Gene-specific short hairpin RNA (shRNA) was synthesized by General Biosystem (Anhui, China), and the target mRNA sequences are 5′-CTTCGACTGTGTCACTATAGG-3′ (Mus musculus CLEC5A, [239]NM_021364) and 5′-GAATCATCTTGGAATGAAAGC-3′ (Homo sapiens CLEC5A, [240]NM_001301167). AAV serotype 9 carrying the endothelial-specific promoter Tie1 (AAV9-Tie1) and lentivirus (LV)–mediated gene transfer was used for in vivo and in vitro studies of CLEC5A, respectively. CLEC5A gene-specific shRNA or coding sequence (General Biosystem) was ligated into pAAV-CMV-U6 (Wanleibio, Shengyang, China), pLVX-shRNA1, or pLVX-IRES-puro (BR004# and BR025#, Fenghui Biotechnology, Hunan, China). For in vitro experiments, endothelial cells were infected with lentivirus expressing CLEC5A (CLEC5A^oe) or shRNA against CLEC5A (CLEC5A^sh). For in vivo experiments, AAV9 expressing CLEC5A or gene-specific shRNA under the promoter of Tie1 (CLEC5A^Tie1-oe or CLEC5A^Tie1-sh) was delivered into mice through tail vein injection ([241]57). Virus carrying empty vector or expressing shRNA that targets a nonspecific sequence was used as the negative control. For in vivo studies of CLEC5A downstream targets, CLEC5A^−/− mice were injected with AAV9-Tie1 expressing specific shRNA targeting SCARB1 or PODXL, or coding sequence of RAMP2 or FABP4 under Tie1 via tail vein. Fourteen days after injection of 100 μl of virus containing 1 × 10^10 AAV9 vector genomes, mice received CLP surgery as described above. The efficiency of AAV/LV-mediated gene transfer in vivo and in vitro was validated by real-time polymerase chain reaction (PCR). Cells and treatment Mouse PMVECs were isolated as previously described ([242]58). Cells were identified as positive immunostaining for CD31 and cultured in primary endothelial cell culture system (PriMed-iCell-002#, iCell, Shanghai, China). Cells between passages 3 and 6 were used for the experiments. Mouse alveolar type II epithelial cells were isolated by magnetic sorting ([243]59) and identified by staining for SFTPC (A23181#, ABclonal, Shanghai, China). Mouse peripheral blood monocytes (PBMCs) were isolated and purified using the commercial kit (P5230#, Solarbio, Beijing, China). Mouse peripheral blood granulocytes were isolated according to a previously described method ([244]60). HUVECs, HMVECs, and THP-1 cells were purchased from iCell Bioscience Inc. (Shanghai, China). HUVECs were cultured in HUVEC culture medium (iCell-h110-001b#, iCell). THP-1 cells were cultured in RPMI 1640 medium (31800#, Solarbio) supplemented with 10% fetal bovine serum and 0.05 mM β-mercaptoethanol. For inflammatory stimulation, mouse PMVECs or HUVECs were exposed to LPS (10 μg/ml) and the duration depended on experimental design. All in vitro experiments were conducted with N = 4 biological repetitions per group. Endothelial cell permeability assay Permeability assays for mouse PMVECs monolayers were carried out following a previous report ([245]61). Cells were seeded onto transwell inserts with 0.4-μm pore size (14312#, LABSELECT, Hefei, Anhui, China) in 24-well plates at a density of 4 × 10^4 per well. The cells, after monolayers were formed, were treated with LPS (10 μg/ml) for 24 hours, followed by incubation with fluorescein isothiocyanate (FITC)–labeled dextran (1:50) for 45 min. The permeability of monolayers was assessed by detecting the fluorescence of the lower chamber medium ([246]62). TEER measurement The barrier integrity and function of mouse PMVECs and HUVECs was assessed in vitro by measuring TEER (Millicell ERS-2#, Millipore, Darmstadt, Germany). Cells were seeded onto 0.1% gelatin-coated transwell inserts (14312#, LABSELECT) in 24-well plates and grown to confluence. After the basal TEER was stabilized, cells were treated with LPS (10 μg/ml) for 4 hours, and the resistance was measured and normalized to the TEER value across an empty insert. Monocyte-endothelial adhesion assay The adhesion of mouse peripheral blood monocytes to PMVECs and human monocytes THP-1 to HUVECs were assessed. Endothelial cells were seeded into 96-well plates and treated with LPS (10 μg/ml) for 5 hours. After removal of LPS, the cells were cocultured with BCECF-AM–labeled monocytes for 1 hour. The adherent cells were harvested and lysed in 0.2% Triton X-100 buffer. The fluorescence intensity was measured, and the data are expressed as the percentage relative to the control group. Trans-endothelial migration assay Trans-endothelial migration of monocytes was performed as previously described ([247]61). Mouse PMVECs were seeded onto collagen I (10 μg/cm^2; BS929-1g#, Biosharp, Hefei, China)–coated transwell inserts (3-μm pore size; 14321#, LABSELECT) in 24-well plates. Cells were cultured until monolayer formation, and then exposed to LPS (10 μg/ml) for 5 hours. After coculturing with mouse peripheral blood monocytes or granulocytes (2 × 10^5 per well) for 4 hours, cells in the lower chamber were collected and cell morphology was examined under a phase contrast microscopy (IX53#, Olympus, Tokyo, Japan). Granulocytes across the endothelial monolayer were identified as 100% neutrophils. The number of transmigrated cells was determined using 3-(4,5-Dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) assay. THP-1 trans-endothelial migration by coculturing with HUVECs was assessed as described above. Western blotting Western blotting was performed according to a standard protocol. Briefly, total protein was extracted with the use of radioimmunoprecipitation assay buffer (R0010#, Solarbio), and an equal amount of protein was loaded and separated by SDS–polyacrylamide gel electrophoresis and transferred onto a polyvinylidene difluoride membrane (IPVH00010#, Millipore). The blot was probed with anti-CLEC5A (PA5-34928#, Thermo Fisher Scientific, Pittsburgh, PA, USA), anti-CD31 (ab222783#, Abcam), anti-VE-cadherin (ab205336#, Abcam), anti-SCARB1 (A1584#, ABclonal Technology, Wuhan, China), PODXL (A10200#), RAMP2 (A3075#), or FABP4 (A11481#) at a dilution of 1:500 or 1:1000, followed by anti-rabbit IgG, horseradish peroxidase–conjugated secondary antibody (SE134#, Solarbio). The target bands were developed using an enhanced chemiluminescent reagent (PE0010#, Solarbio), and analyzed by Gel-Pro-Analyzer software. The immunoblot was stripped and reprobed with glyceraldehyde-3-phosphate dehydrogenase (GAPDH) (60004-1-Ig#, Proteintech Group Inc., Rosemont, IL, USA), which serves as an internal reference. Quantitative real-time PCR Total RNA was extracted using TRIpure kit (RP1001#, BioTeke, Beijing, China) according to the manufacturer’s instructions. RNA concentration was determined, and reverse transcription was performed using the BeyoRT II M-MLV reverse transcriptase (D7160L#, Beyotime, Shanghai, China). The quantitative PCR was carried out with 2×Taq PCR MasterMix (PC1150#, Solarbio) and SYBR Green (SY1020#, Solarbio) in Exicycler 96 (Bioneer, Daejeon, Korea). Sequences of specific PCR primers were included in table S1. The relative quantification was calculated using the 2^−ΔΔCT method, and the data are displayed relative to the control group that arbitrarily set as 100%. Enzyme-linked immunosorbent assay Mouse PMVECs were exposed to LPS (10 μg/ml) for 24 hours, and the supernatant was collected for cytokine detection. Levels of TNF-α, IL-6, MCP-1, ICAM-1, and VCAM-1 were determined by commercial enzyme-linked immunosorbent assay (ELISA) kits from Lianke Technology (EK282#, EK206#, EK287#, EK289#, and EK290#, Hangzhou, China) and CXCL5 by FineTest (EM1004#, Wuhan, China). Statistical analysis All quantification in this study was carried out in a blinded manner. No sample or data were excluded from analysis. All figures shown are representative images that reflected the average level of at least three independent experiments. All data are analyzed using GraphPad Prism 8.0 software, and the graphs are shown as mean (SD) values of six biological repetitions for in vivo experiments and four biological repetitions for in vitro experiments. Data that meet the normal distribution by the Shapiro-Wilk test were subjected to variance analysis. For comparisons among three or more groups, the homogeneity of variance was confirmed by the Brown-Forsythe test or Bartlett’s test, and one-way analysis of variance (ANOVA) analysis or Brown-Forsythe and Welch ANOVA test was performed. Comparisons between groups were carried out by Tukey’s multiple comparisons test or Games-Howell’s multiple comparisons test. Lung injury score was analyzed by the Kruskal-Wallis test followed by Dunn’s multiple comparisons test. An unpaired two-tailed t test was used for comparisons between two groups with F test to compare variances or Mann-Whitney test. Data of CLEC5A expression at different time points following LPS or CLP challenge were performed with repeated measures ANOVA analysis with the Geisser-Greenhouse correction, followed by Dunnett’s or Sidak’s multiple comparisons test. For survival analysis, the log-rank (Mantel-Cox) test, log-rank for trend, or Gehan-Breslow-Wilcoxon test was used for comparisons of survival curves between groups. A P value less than 0.05 was considered statistically significant. Acknowledgments