Abstract
   Tumor-derived exosomes carry programmed death-ligand 1 (PD-L1), which
   binds programmed cell death protein 1 (PD-1) on T cells, suppressing
   immune responses locally and systemically. However, the mechanisms
   governing exosomal PD-L1 sorting and secretion remain elusive. Here, we
   identify Munc13-4 as a crucial regulator of this process. Deletion of
   Munc13-4 in breast tumors enhances T cell-mediated anti-tumor immunity,
   suppresses tumor growth, and improves the efficacy of immune checkpoint
   inhibitors. Mechanistically, Munc13-4 collaborates with hepatocyte
   growth factor-regulated tyrosine kinase substrate (HRS), Rab27, and
   SNAREs to facilitate PD-L1 sorting and secretion via exosomes.
   Cryogenic electron microscopy (cryo-EM) analysis of the Munc13-4–Rab27a
   complex provide structural insights into exosome secretion.
   Importantly, PD-L1 sorting relies on a ternary complex composed of
   Munc13-4, PD-L1 and HRS, which is regulated by interferon gamma (IFNγ)
   signaling. A designed peptide that disrupts Munc13-4–PD-L1 interaction
   impedes PD-L1 sorting, enhances antitumor immunity, and suppresses
   tumor growth, highlighting the therapeutic potential of targeting this
   pathway.
   Subject terms: Membrane trafficking, Cancer
     __________________________________________________________________
   Munc13-4 is known to regulate a Ca^2+- stimulated exosome release
   pathway. Here, the authors discover that Munc13-4 promotes breast
   cancer immune evasion through regulation of PD-L1 sorting in exosomes.
Introduction
   Tumor cells evade immune surveillance by increasing the surface
   expression of PD-L1, which interacts with the PD-1 receptor on T cells
   to trigger the immune checkpoint response and leading to T cell
   inhibition^[44]1,[45]2. Inhibitors targeting the PD-1/PD-L1 have shown
   promise in cancer therapy by restoring T cell function and enhancing
   anti-tumor immunity^[46]3. Nonetheless, a significant proportion of
   patients do not respond to anti-PD-L1/PD-1 therapies^[47]3,[48]4. A
   primary reason for this resistance is the secretion of PD-L1 into the
   bloodstream via extracellular vesicles (EVs), particular exosomes,
   where it can disrupt immune function remotely^[49]5–[50]7. Recent
   studies have demonstrated that genetic blockade of exosomal PD-L1
   biogenesis and/or secretion not only suppresses local tumor growth but
   also elicits a durable systemic anti-tumor immune response^[51]8.
   Therefore, the role of exosomal PD-L1 in immune modulation highlights
   the urgent need for targeted strategies to inhibit exosomal PD-L1
   biogenesis and secretion, which could significantly improve the
   efficacy of cancer immunotherapies.
   The secretion of tumor-derived exosomes begins with the formation of
   early endosomes, which mature into multivesicular bodies (MVBs)
   containing intraluminal vesicles (ILVs). MVBs fuse with the plasma
   membrane, releasing ILVs as exosomes into the extracellular
   space^[52]9. Endosomal sorting complexes required for transport
   (ESCRTs) are critical for cargo sorting within the endosomal system.
   HRS, a key component of the ESCRT machinery, mediates cargo recognition
   and sorting into MVBs^[53]10. In addition, the Rab family of small
   GTPases (Rabs) is essential for membrane trafficking, with Rab27
   specifically controlling various steps of the exosome secretion
   pathway, particularly the docking of MVBs to the plasma
   membrane^[54]11. Moreover, the SNARE machinery mediates the fusion of
   MVBs with the plasma membrane, with syntaxin-4, SNAP-23, and VAMP-7
   forming the SNARE complex to drive the secretion of exosomes in various
   tumor cells^[55]12. Despite these advances, the mechanisms governing
   PD-L1 sorting onto exosomes and their secretion remain elusive.
   Understanding these processes is essential for developing targeted
   molecules that inhibit the PD-L1 secretion pathway, and it is urgent to
   identify and investigate key players in the endosomal sorting pathway
   that mediate PD-L1 recognition and sorting onto exosomes.
   In this study, we identified Munc13-4, a member of the Munc13 protein
   family known for its regulatory role in membrane trafficking, as
   upregulated in various tumor tissues, where it mediates tumor immune
   evasion by regulating exosomal sorting and secretion of PD-L1. Deleting
   Munc13-4 in breast tumors significantly enhances T cell-mediated
   anti-tumor immunity, suppresses tumor growth, and boosts the efficacy
   of immune checkpoint inhibitors. We elucidated a coherent mechanism
   whereby Munc13-4 collaborates with HRS, Rab27, and SNAREs to regulate
   PD-L1 sorting and secretion via exosomes. Notably, we discovered a
   critical function of Munc13-4 in PD-L1 sorting that depends on its
   direct interaction with PD-L1; disrupting this interaction with a
   specifically designed peptide markedly impaired PD-L1 sorting, leading
   to enhanced T cell-mediated anti-tumor responses in vivo. These
   findings position Munc13-4 as a promising therapeutic target for
   boosting immune responses against tumors.
Results
Munc13-4 deficiency in tumor cells inhibits tumor growth in an
immunity-dependent way
   The Munc13 protein family functions as critical regulators of vesicle
   trafficking and exocytosis across various cell types. While Munc13-1,
   Munc13-2, and Munc13-3 are involved in the exocytosis of synaptic
   vesicles and dense-core vesicles in neurons and neuroendocrine
   cells^[56]13–[57]15, Munc13-4 has specialized roles in cytotoxic
   granule exocytosis in immune cells^[58]16–[59]19 and has recently been
   implicated in exosome secretion in tumor cells^[60]20. To explore the
   role of Munc13-4 in tumor progression, we assessed its expression in
   tumor and adjacent normal tissues using TIMER2.0
   (cistrome.shinyapps.io/timer)^[61]21,[62]22, a platform that analyzes
   genomic data from The Cancer Genome Atlas (TCGA). Our analysis revealed
   significant upregulation of Munc13-4 across various tumor types
   (Supplementary Fig. [63]1A). Immunohistochemical staining of tumor and
   adjacent normal tissues on tissue microarrays demonstrated increased
   Munc13-4 expression in tumors, including breast cancer, thyroid cancer,
   cholangiocarcinoma, gastrointestinal stromal tumor, pancreatic cancer,
   and hepatocellular carcinoma (Fig. [64]1A and Supplementary
   Fig. [65]1B–E), suggesting a crucial role for Munc13-4 in tumor
   progression. Given the high global incidence of breast cancer, we
   focused on investigating the specific role of Munc13-4 in breast
   cancer.
Fig. 1. Munc13-4 deficiency in tumor cells inhibits tumor growth in an
immunity-dependent way.
   [66]Fig. 1
   [67]Open in a new tab
   A Representative immunohistochemical images showing Munc13-4 expression
   in breast cancer (n = 30 patients), thyroid cancer (n = 6 patients),
   cholangiocarcinoma (n = 5 patients), gastrointestinal stromal tumors
   (n = 6 patients), pancreatic cancer (n = 6 patients), and
   hepatocellular carcinoma (n = 9 patients) tissues, along with their
   corresponding adjacent normal tissues, assessed using a multi-organ
   carcinoma tissue array. Scale bar, 500 μm. B–D Tumor growth in BALB/c
   mice inoculated with wild-type (WT), control, or Munc13-4 knockout (KO)
   4T1 cells (n = 9 mice). B Schematic of experimental design. C Tumor
   growth curves following mammary gland inoculation. D Percentage change
   in tumor volume, normalized to WT group. E–G Tumor growth in BALB/Nude
   mice inoculated with WT, control, or Munc13-4 KO 4T1 cells (n = 8
   mice). E Schematic of experimental design. F Tumor growth curves
   following mammary gland inoculation. G Percentage change in tumor
   volume, normalized to WT group. H–J Tumor growth in NOD/SCID mice
   inoculated with WT, control, or Munc13-4 KO 4T1 cells (n = 6 mice). H
   Schematic of experimental design. I Tumor growth curves following
   mammary gland inoculation. J Percentage change in tumor volume,
   normalized to WT group. Data are presented as means ± SEM, and p-values
   were calculated by one-way ANOVA with multiple comparisons (C, F, I),
   ns not significant. Source data are provided as a [68]Source data file.
   Using the CRISPR-Cas9 system, we generated Munc13-4 knockout 4T1 murine
   mammary carcinoma cells, with control cells infected with a lentivirus
   carrying Cas9 without sgRNA (Supplementary Fig. [69]2A). The deletion
   of Munc13-4 did not influence the proliferation of 4T1 cells in vitro
   (Supplementary Fig. [70]2B). We then conducted in vivo studies by
   creating orthotopic mouse models of breast cancer using wild-type (WT),
   control and Munc13-4 knockout 4T1 cells (Fig. [71]1B). Mice inoculated
   with Munc13-4 knockout 4T1 cells showed a significant delay in tumor
   growth compared to those implanted with WT or control cells
   (Fig. [72]1C and Supplementary Fig. [73]2C, D), with average tumor
   volume in the knockout group stabilizing at approximately 17.17% of
   that in the WT group (Fig. [74]1D). These data demonstrate that
   Munc13-4 knockout substantially impairs the oncogenic potential of 4T1
   cells, indicating a pivotal role for Munc13-4 in breast tumor
   progression.
   Next, we identified differentially expressed proteins in the proteomes
   of control and Munc13-4 knockout 4T1 cells, and conducted enrichment
   analysis based on the KEGG database to pinpoint relevant cellular
   processes and organismal systems affected by Munc13-4 knockout. KEGG
   enrichment analysis suggests significant involvement in transport and
   catabolic pathways, along with a critical association with the immune
   system (Supplementary Fig. [75]2E). Given that immune evasion is a
   hallmark of cancer, we explored the relationship between Munc13-4 and
   immune evasion by assessing the oncogenicity of Munc13-4 knockout 4T1
   cells in immunodeficient mouse models. In BALB/Nude mice lacking T
   cells, the difference in tumor growth between those with Munc13-4
   knockout cells and WT or control cells became less pronounced
   (Fig. [76]1E, F and Supplementary Fig. [77]2F, G), with Munc13-4
   knockout tumors approaching 60.19% of the size of WT tumors
   (Fig. [78]1G). Moreover, in NOD/SCID mice with severe combined
   immunodeficiency, tumors in the Munc13-4 knockout group grew comparably
   to those in the WT group (Fig. [79]1H, I and Supplementary Fig. [80]2H,
   I), with the Munc13-4 knockout tumors reaching 83.67% of the size of
   the WT tumors (Fig. [81]1J). These results suggest that the role of
   Munc13-4 in breast tumor progression is closely linked to its capacity
   to modulate immune responses within the tumor microenvironment.
Munc13-4 deficiency in tumor cells enhances T cell infiltration and
activation
   We next explored whether deletion of Munc13-4 in breast tumor cells
   influences the quantity and activity of T cells within tumors, spleens,
   and lymph nodes of tumor-bearing mice inoculated with either control or
   Munc13-4 knockout 4T1 cells. Flow cytometry showed a significant
   increase in the infiltration of both CD4^+ and CD8^+ T cells within the
   tumors of mice implanted with Munc13-4 knockout 4T1 cells (Fig. [82]2A
   and Supplementary Fig. [83]3A), which was corroborated by
   immunofluorescent staining (Supplementary Fig. [84]3B). In addition,
   the populations of both CD4^+ and CD8^+ cells in the spleens and lymph
   nodes of these mice were notably elevated compared to the control group
   (Fig. [85]2B, C). These results indicate that the deletion of Munc13-4
   in tumor cells enhances T cell infiltration in the tumor, spleen, and
   lymph nodes.
Fig. 2. Munc13-4 deficiency in tumor cells enhances T cell infiltration and
activation.
   [86]Fig. 2
   [87]Open in a new tab
   Flow cytometric quantification of the percentage of CD45^+CD3^+CD4^+
   and CD45^+CD3^+CD8^+ T cells among total cells in the tumors (A),
   spleens (B), and draining lymph nodes (C) of BALB/c mice (n = 5 mice),
   21 days after mammary gland injection with 3 × 10^5 control or Munc13-4
   KO 4T1 cells per mouse. Quantification of the percentage of granzyme
   B^+ (GzmB^+) (D), Ki67^+ (E), and IFNγ^+ (F) cells among
   CD45^+CD3^+CD4^+ and CD45^+CD3^+CD8^+ T cells within tumors from
   orthotopic mouse models of breast cancer generated by control or
   Munc13-4 KO 4T1 cells (n = 5 mice). Quantification of the percentage of
   granzyme B^+ (G), Ki67^+ (H), and IFNγ^+ (I) cells among
   CD45^+CD3^+CD4^+ and CD45^+CD3^+CD8^+ T cells within spleens from
   orthotopic mouse models of breast cancer generated by control or
   Munc13-4 KO 4T1 cells (n = 5 mice). Quantification of the percentage of
   granzyme B^+ (J), Ki67^+ (K), and IFNγ^+ (L) cells among
   CD45^+CD3^+CD4^+ and CD45^+CD3^+CD8^+ T cells within the draining lymph
   nodes from orthotopic mouse models of breast cancer generated by
   control or Munc13-4 KO 4T1 cells (n = 5 mice). Box plots show the
   median (center line), interquartile range (box), minima and maxima
   (whiskers), and all individual data points (dots). All p-values were
   calculated by two-tailed Multiple t-tests. Source data are provided as
   a [88]Source data file.
   We further assessed T cell activation markers, including the cytotoxic
   molecule granzyme B, the proliferation marker Ki67, and the cytokine
   interferon gamma (IFNγ), in CD4^+ and CD8^+ T cell populations from
   tumors, spleens, and lymph nodes of tumor-bearing mice. Mice implanted
   with Munc13-4 knockout 4T1 cells exhibited a significant increase in
   the expression of granzyme B, Ki67, and IFNγ in both CD4^+ and CD8^+ T
   cells across all examined tissues compared to those inoculated with
   control 4T1 cells (Fig. [89]2D–L and Supplementary Fig. [90]3C–E).
   Collectively, these results demonstrate that Munc13-4 deficiency in
   breast tumor cells augments T cell infiltration and activation, thereby
   enhancing the systemic T cell-mediated immune response.
Facilitating PD-L1 secretion by Munc13-4 suppresses anti-tumor efficacy of T
cells
   PD-L1 binding to PD-1 on T cells is a crucial mechanism for tumor
   evasion of immune surveillance. To explore whether reduced
   tumor-induced immunosuppression in Munc13-4-deficient models was linked
   to PD-L1 changes, we examined the effect of Munc13-4 knockout on PD-L1
   expression by western blot. Our results showed no impact on PD-L1
   protein levels in 4T1 and SUM159 breast tumor cells (Fig. [91]3A).
Fig. 3. Facilitating PD-L1 secretion by Munc13-4 suppresses anti-tumor
efficacy of T cells.
   [92]Fig. 3
   [93]Open in a new tab
   A Western blot (WB) analysis of total PD-L1 level in control and
   Munc13-4 KO SUM159 or 4T1 cells (n = 3 biological replicates). B
   Representative TEM images of EVs secreted by control and Munc13-4 KO
   SUM159 or 4T1 cells (n = 3 biological replicates). Scale bar, 50 nm. C
   Quantification of exosomes secreted by equal numbers of control and
   Munc13-4 KO SUM159 (left) or 4T1 (right) cells through NTA (n = 3
   biological replicates). D Schematic of experimental design for (E, F).
   WB analysis of PD-L1, Alix, CD63, and CD81 in EVs secreted by equal
   numbers of control and Munc13-4 KO SUM159 (E) or 4T1 (F) cell,
   collected from factions 1–6 (F1–6) (n = 3 biological replicates). WB
   analysis of PD-L1, Alix, CD63, and CD81 abundance on equal numbers of
   exosomes secreted by control and Munc13-4 KO SUM159 cells (G) and
   corresponding quantification of blot band intensities (H) (n = 4
   biological replicates). WB analysis of PD-L1, Alix, CD63, and CD81
   abundance on equal numbers of exosomes secreted by control and Munc13-4
   KO 4T1 cells (I) and corresponding quantification of blot band
   intensities (J) (n = 3 biological replicates). K Schematic of
   experimental design for (L). L Quantification of killing efficiency
   against control and Munc13-4 KO 4T1 cells (n = 3 biological
   replicates). M Schematic of experimental design for (N). N Tumor growth
   curves following mammary gland inoculation of control or Munc13-4 KO
   4T1 cells, with subsequent injection of PBS or the indicated exosomes
   (n = 6 mice). In (A), for each cell line, all samples were run on the
   same gel; in (E, F, G, I), samples were run on the same gel. Data are
   presented as means ± SEM (C, H, J, N), and p-values were calculated by
   unpaired two-tailed t-test (C), two-way ANOVA (H, J) with multiple
   comparisons, paired two-tailed t-test (L), and one-way ANOVA with
   multiple comparisons (N). Source data are provided as a [94]Source data
   file.
   PD-L1 transport to the plasma membrane and sorting onto exosomes
   enables its interaction with PD-1 on T cells, thus inhibiting T
   cell-mediated anti-tumor immunity. Proteomics analysis suggests the
   involvement of Munc13-4 in transport and catabolic pathways
   (Supplementary Fig. [95]2E), we thus examined the effect of Munc13-4
   knockout on the transport of PD-L1 to the plasma membrane and to
   exosomes. Flow cytometry analyses revealed that no change in PD-L1 on
   the plasma membrane was observed after Munc13-4 knockout (Supplementary
   Fig. [96]4A, B). Transmission electron microscopy (TEM), nanoparticle
   tracking analysis (NTA) and tunable resistive pulse sensing (TRPS)
   showed that EVs isolated from control and Munc13-4 knockout cell
   culture supernatants displayed exosome-like morphology (Fig. [97]3B)
   and were within the 50–200 nm size range characteristic of exosomes
   (Supplementary Fig. [98]4C–F). However, NTA and TRPS indicated a
   significant decrease in the total number of exosomes secreted by
   Munc13-4 knockout cells compared to control cells (Fig. [99]3C and
   Supplementary Fig. [100]4C–F). Western analysis showed a marked
   reduction in PD-L1 and exosome marker proteins (Alix, CD63, CD81) in
   EVs from Munc13-4 knockout cells (Supplementary Fig. [101]4G, H).
   Isolation of EVs via Optiprep^TM density gradient centrifugation
   confirmed the association of PD-L1 with exosomes and indicated
   disrupted exosome secretion in Munc13-4 knockout cells
   (Fig. [102]3D–F).
   To assess PD-L1 levels on exosomes, we collected equivalent numbers of
   exosomes from both control and knockout cells. Western analysis showed
   that PD-L1 levels were significantly lower in exosomes from Munc13-4
   knockout cells, while levels of Alix, CD63, and CD81 remained unchanged
   (Fig. [103]3G–J). These results indicate that Munc13-4 deletion does
   not affect the total PD-L1 levels or its presence on the plasma
   membrane but significantly reduces secreted PD-L1 by inhibiting
   exosomes secretion and decreasing its enrichment onto exosomes.
   We further assessed the role of exosomes from control and Munc13-4
   knockout cells in T cell suppression by flow cytometry. Again,
   equivalent numbers of exosomes from both control and knockout cells
   were collected. Exosomes from Munc13-4-deficient 4T1 cells showed
   reduced inhibitory effects on the cytotoxicity of CD8^+ T cells
   compared to exosomes from control cells (Supplementary Fig. [104]4I,
   J). Notably, anti-PD-L1 treatment significantly decreased the
   suppression of CD8^+ T cell activation by control exosomes, while it
   had minimal effect on exosomes from Munc13-4 knockout cells
   (Supplementary Fig. [105]4I, J). Taken together, the reduced PD-L1
   presence on exosomes due to Munc13-4 knockout in tumor cells leads to
   decreased immunosuppressive capacity.
   To directly characterize the anti-tumor efficacy of T cells influenced
   by Munc13-4 deficiency in tumor cells, we examined the cytotoxicity of
   T cells primed by either control or Munc13-4-deficient 4T1 cells in
   vitro. Mouse spleen lymphocytes activated with anti-CD3 and anti-CD28
   antibodies displayed significantly enhanced cytotoxicity against
   Munc13-4 knockout 4T1 cells compared to control cells (Fig. [106]3K,
   L). Meanwhile, we explored the relationship between decreased
   oncogenicity and impaired exosomal PD-L1 secretion in
   Munc13-4-deficient 4T1 cells in vivo (Fig. [107]3M). Infusion of
   exosomes from control 4T1 cells markedly accelerated tumor growth in
   mice bearing Munc13-4 knockout 4T1 cell transplants (Fig. [108]3N and
   Supplementary Fig. [109]4K, L). In contrast, exosomes pre-treated with
   anti-PD-L1 antibody had minimal impact on tumor growth (Fig. [110]3N
   and Supplementary Fig. [111]4K, L). Collectively, these findings
   underscore the essential role of Munc13-4 in T cell suppression and
   tumor progression through its regulation of PD-L1 secretion via
   exosomes.
Munc13-4 deficiency in tumor cells boosts immune checkpoint blockade therapy
effectiveness
   Immune checkpoint blockade (ICB) therapy, which targets the PD-1/PD-L1
   interaction using antibodies, has become a common approach in cancer
   treatment. However, it faces challenges such as limited durability of
   remission and a low overall response rate, restricting its benefits to
   a small subset of patients^[112]3,[113]4. Recent studies indicate that
   PD-L1 on exosomes secreted by tumor cells may antagonize ICB
   therapy^[114]5–[115]7. Given our findings that Munc13-4 promotes PD-L1
   secretion and inhibits immune surveillance in vivo, we investigated
   whether Munc13-4 deletion in tumor cells could enhance ICB therapy
   efficacy (Supplementary Fig. [116]5A).
   In experiments, neither anti-PD-1 nor anti-PD-L1 treatment slowed tumor
   growth in mice inoculated with control 4T1 cells compared to IgG
   isotype controls (Supplementary Fig. [117]5B–D). In contrast, tumor
   growth was significantly delayed in mice implanted with
   Munc13-4-deficient 4T1 cells, and this effect was further enhanced by
   anti-PD-1 and anti-PD-L1 treatments (Supplementary Fig. [118]5B–D).
   These results suggest that Munc13-4 depletion in tumor cells improves
   the therapeutic efficacy of immune checkpoint inhibitors.
Munc13-4 does not influence MVB biogenesis
   The above findings indicate that Munc13-4 deletion in breast tumor
   cells has two detrimental effects on PD-L1 secretion: (i) reduced PD-L1
   sorting on exosomes and (ii) impaired exosome secretion. This may
   result from impaired MVB biogenesis and/or MVB fusion with the plasma
   membrane. To explore the underlying mechanisms, we utilized the human
   breast tumor cell line SUM159 and examined whether Munc13-4 deficiency
   disrupts MVB biogenesis. TEM analysis of control and Munc13-4 knockout
   SUM159 cells revealed a significant accumulation of MVBs in the
   knockout cells (Supplementary Fig. [119]6A, B). This was corroborated
   by immunofluorescence and western analyses showing a marked increase in
   CD63, a well-established MVB marker (Supplementary Fig. [120]6F–H). In
   addition, TEM analysis indicated an increase in ILVs within MVBs in
   Munc13-4 knockout cells (Supplementary Fig. [121]6C), suggesting that
   reduced PD-L1 secretion is not due to impaired MVB biogenesis.
   Further TEM analysis showed a significant increase in hybrid structures
   formed between MVBs and lysosomes in Munc13-4 knockout cells,
   characterized by electron-dense compartments and double-membrane
   autophagosomes (Supplementary Fig. [122]6A, D). Immunofluorescence
   analysis confirmed increased colocalization of CD63 with the lysosomal
   marker LAMP1 (Supplementary Fig. [123]6I, J). Moreover, TEM data
   indicated that the diameter of MVBs in Munc13-4 knockout cells was
   significantly enlarged (Supplementary Fig. [124]6E). Together, these
   observations suggest an increased prevalence of both homotypic fusion
   among MVBs and heterotypic fusion between MVBs and lysosomes in the
   absence of Munc13-4.
Munc13-4 facilitates MVB docking and fusion with the plasma membrane
   Since Munc13-4 deletion does not affect MVB biogenesis, we investigated
   its role in downstream processes related to the docking and fusion of
   MVBs with the plasma membrane. Rab GTPase Rab27a plays a critical role
   in various stages of the exosome secretion pathway, particularly in the
   transport and docking of MVBs to the plasma membrane^[125]11. As an
   effector of Rab27a, Munc13-4 collaborates with Rab27a to regulate
   exocytosis in various immune cells^[126]19,[127]23,[128]24. Consistent
   with this role, our in vitro binding experiment detected significant
   binding between Munc13-4 and Rab27a (Supplementary Fig. [129]6K). In
   contrast, only minimal interactions were detected with Rab5 or Rab7
   (Supplementary Fig. [130]6K), which are primarily associated with early
   and late endosomes, respectively. To study the role of the
   Munc13-4–Rab27a complex in MVB docking, we expressed CD63 tagged with
   orange fluorescent protein in both control and Munc13-4 (or Rab27a)
   knockout SUM159 cells. By using total internal reflection fluorescence
   (TIRF) microscopy, we tracked the movement of MVBs near the plasma
   membrane^[131]11. Either Munc13-4 or Rab27a deficiency increased MVB
   mobility (Supplementary Fig. [132]6L, M), indicating the functional
   significance of Munc13-4 and Rab27a in MVB docking.
   To investigate the molecular mechanisms by which Munc13-4 and Rab27a
   contribute to MVB docking, we determined the cryo-EM structure of the
   Munc13-4–Rab27a complex (Fig. [133]4A and Supplementary Fig. [134]7).
   Note that the non-hydrolyzable GTP analog GppNHp was added to maintain
   Rab27a in its active state for efficient binding to Munc13-4
   (Supplementary Fig. [135]8A–C). The structure was resolved to 3.4 Å in
   the core region of the complex, while the two C[2] domains at the N-
   and C-termini of Munc13-4 exhibited a resolution of 4.4 Å due to their
   inherent flexibility (Supplementary Fig. [136]8D–H). Compared to the
   solved structure of core domain (C[1]-C[2]B-MUN) of Munc13-1^[137]25,
   the overall architecture of Munc13-4 is more curved (Supplementary
   Fig. [138]8I). The binding interface of the complex comprises residues
   F46, W73, and F88 on Rab27a, along with N739, T740, V660, and K661 on
   Munc13-4 (Fig. [139]4A). Mutations at these sites significantly
   impaired the interaction (Fig. [140]4B, C and Supplementary
   Fig. [141]8J, K). Consistent with these findings, such mutations
   resulted in increased MVB mobility (Fig. [142]4D, E) and a marked
   decrease in the total number of exosomes (Fig. [143]4F). Hence, the
   structure information provides mechanistic insight into how Munc13-4
   works with Rab27a to promote MVB docking.
Fig. 4. Munc13-4 facilitates MVB docking and fusion with the plasma membrane.
   [144]Fig. 4
   [145]Open in a new tab
   A Cryo-EM structure of the Munc13-4–Rab27a complex (upper panel) and
   detailed interface view (lower panel). The corresponding cryo-EM
   density map is shown as a semi-transparent surface. GST pull-down
   assays examining the effects of different mutations in Munc13-4 (B),
   and diverse mutations in Rab27a (C), on the formation of the
   Munc13-4–Rab27a complex (n = 3 biological replicates). Quantification
   of the mean diffusion coefficient (D) detected by TIRF microscopy in
   SUM159 cells with the indicated mutations in Munc13-4 (D) or Rab27a (E)
   (n = 3 biological replicates). F Quantification of exosomes secreted by
   equal numbers of indicated SUM159 cells through NTA (n = 3 biological
   replicates). G Illustration of FRET assay for SNARE complex assembly
   detection. VAMP-7 SNARE motif (V7 SNARE) labeled with donor dye BODIPY
   FL, SNAP-23 (SN-23) labeled with acceptor dye 5-TAMRA, and syntaxin-4
   (transmembrane domain deleted, Syx-4 ΔTM) form a SNARE complex leading
   to FRET. H Representative graph of time-dependent SNARE complex
   assembly measured by the development of FRET (n = 3 biological
   replicates). I Illustration of the liposome fusion experiment.
   Syntaxin-4 (Syx-4) was incorporated into DiD-labeled liposomes, and
   VAMP-7 was incorporated into DiI-labeled liposomes. Munc13-4
   accelerates liposome fusion mediated by SNARE complex, leading to FRET.
   J Time-dependent liposome fusion measured from the development of FRET
   (n = 3 biological replicates). K Quantification of the FRET efficiency
   at the end of the detection (n = 3 biological replicates). Data are
   represented as means ± SEM (D, E, F, J, K), and p-values were
   calculated by one-way ANOVA with multiple comparisons (D, E, F, K).
   Source data are provided as a [146]Source data file.
   Our previous studies have identified a SNARE complex composed of
   syntaxin-4, SNAP-23, and VAMP-7 as mediators of MVB fusion with the
   plasma membrane for exosome release in various tumor cells^[147]12. Our
   binding assays demonstrated a direct interaction between Munc13-4 and
   both SNAP-23 and VAMP-7, and with the assembled SNARE complex
   (Supplementary Fig. [148]6N). We further assessed the regulatory role
   of Munc13-4 in SNARE complex assembly and membrane fusion using
   FRET-based assembly and fusion assays. Our results indicated that
   Munc13-4 significantly facilitates the assembly of syntaxin-4, SNAP-23,
   and VAMP-7 into the SNARE complex (Fig. [149]4G, H) and promotes the
   fusion between liposomes bearing syntaxin-4/SNAP-23 and liposomes
   containing VAMP-7 (Fig. [150]4I–K), underscoring its critical role in
   the fusion of MVBs with the plasma membrane. Altogether, these results
   suggest that Munc13-4 works with Rab27a and SNAREs to complete exosome
   secretion by facilitating the docking and fusion of MVBs with the
   plasma membrane.
Exosomal sorting of PD-L1 by Munc13-4 and HRS
   Consistent with our finding that Munc13-4 knockout reduces PD-L1
   abundance on exosomes, PD-L1 showed decreased colocalization with CD63
   and increased colocalization with LAMP1 in Munc13-4 knockout SUM159
   cells (Fig. [151]5A, B), indicating improper translocation of PD-L1
   from MVBs to lysosomes and strengthening the role of Munc13-4 in
   sorting PD-L1 to exosomes.
Fig. 5. Exosomal sorting of PD-L1 by Munc13-4 and HRS.
   [152]Fig. 5
   [153]Open in a new tab
   A Representative confocal images of SUM159 cells co-expressing
   GFP-PD-L1 with Orange-CD63 or Orange-LAMP1 (3 independent experiments).
   Scale bar, 20 μm. B Quantification of the mean Pearson’s correlation
   coefficient between PD-L1 and CD63 or LAMP1 (n = 3 independent
   experiments). C WB analysis of PD-L1, Alix, CD63, and CD81 on equal
   numbers of exosomes from SUM159 cells (n = 2 biological replicates). D
   WB analysis of total HRS in SUM159 cells (n = 3 biological replicates).
   E Co-IP/immunoblotting (IB) analysis in SUM159 cells transfected with
   indicated constructs (n = 3 biological replicates). F Representative
   confocal images of SUM159 cells in the PLA (3 independent experiments).
   Scale bar, 10 μm. G Quantification of the mean number of PLA puncta per
   cell (n = 3 independent experiments). H Co-IP/IB analysis in Munc13-4
   KO SUM159 cells transfected with indicated constructs (n = 3 biological
   replicates). I Co-IP/IB analysis in HEK293T cells transfected with
   indicated constructs (n = 3 biological replicates). J Representative
   PLA image of SUM159 cells (3 independent experiments). Scale bar,
   10 μm. K Co-IP/IB analysis in HEK293T cells transfected with indicated
   constructs (n = 3 biological replicates). L Representative PLA image of
   SUM159 cells (3 independent experiments). Scale bar, 10 μm. M Schematic
   of in vitro liposome co-flotation assay. N WB analysis of Munc13-4 and
   PD-L1 in top three fractions and bottom fraction (n = 3 biological
   replicates). O Co-IP/IB analysis in different SUM159 cells transfected
   with indicated constructs (n = 3 biological replicates). In (C, D, N),
   samples were run on the same gel; in (E, H, I, K, O), IP and Input
   samples were derived from the corresponding same experiment, but
   different gels for IP and Input samples were processed in parallel.
   Data are represented as means ± SEM (B, G), and p-values were
   calculated by two-way ANOVA with multiple comparisons (B) and unpaired
   two-tailed t-test (G). Source data are provided as a [154]Source data
   file.
   HRS is a key component of ESCRT-0 and has been found to mediate PD-L1
   sorting^[155]26,[156]27. This function was verified by the significant
   reduction in PD-L1 abundance on exosomes following the deletion of HRS
   (Fig. [157]5C). To further explore this, we examined the relationship
   between HRS and PD-L1 in both the presence and absence of Munc13-4. In
   Munc13-4 knockout SUM159 cells, the expression of PD-L1 and HRS was
   unaffected (Figs. [158]3A and [159]5D); while the interaction between
   HRS and PD-L1 was markedly diminished compared to control cells,
   regardless of whether they were exogenously or endogenously expressed
   (Fig. [160]5E–G). Reintroducing Munc13-4 into knockout cells
   significantly restored the HRS–PD-L1 interaction (Fig. [161]5H),
   suggesting that this interaction relies on Munc13-4. Supporting these
   findings, co-immunoprecipitation (co-IP) assay showed a marked binding
   preference of Munc13-4 for HRS over STAM (another ESCRT-0 component)
   under exogenous expression conditions (Fig. [162]5I). In addition,
   proximity ligation assay (PLA) indicated an interaction between
   endogenous expressed Munc13-4 and HRS in SUM159 cells (Fig. [163]5J).
   We also examined whether Munc13-4 directly interacts with PD-L1. Both
   co-IP and PLA assays revealed this interaction in SUM159 cells
   (Fig. [164]5K, L). This interaction was further verified through in
   vitro liposome co-flotation assay (Fig. [165]5M, N). Importantly, the
   deletion of HRS in SUM159 cells did not influence the interaction
   between Munc13-4 and PD-L1 (Fig. [166]5O), indicating that the
   Munc13-4–PD-L1 interaction is independent of HRS. Altogether, these
   results suggest the formation of a ternary complex comprising HRS,
   Munc13-4, and PD-L1, which is critical for sorting of PD-L1 to
   exosomes.
IFNγ-induced modifications of Munc13-4 and HRS exert opposing effects on
PD-L1 sorting
   IFNγ, a cytokine produced by NK and T cells, contributes substantially
   to immunosurveillance against tumors by activating immune cells and
   inducing apoptosis in tumor cells^[167]28,[168]29. Conversely, tumor
   cells can exploit IFNγ signaling to evade immune destruction through
   the elevation of PD-L1 expression, which inhibits immune cell
   activity^[169]28,[170]29. Indeed, IFNγ stimulation dramatically
   increased the protein level of PD-L1 (Supplementary Fig. [171]9A, B),
   leading to an elevated presence of PD-L1 on both the plasma membrane
   (Supplementary Fig. [172]9C, D) and exosomes (Supplementary
   Fig. [173]9E, F), without altering the total quantity of secreted
   exosomes (Supplementary Fig. [174]9E–G). Considering the significant
   effect of exosomal PD-L1 on the suppression of T cell activity, we
   hence explored whether and how HRS and Munc13-4 regulate PD-L1 sorting
   onto exosomes in response to IFNγ.
   Unlike PD-L1, the overall levels of HRS and Munc13-4 remained unchanged
   under IFNγ treatment in SUM159 cells (Supplementary Fig. [175]9A, B).
   Strikingly, IFNγ stimulation significantly decreased the ubiquitylation
   of HRS (Supplementary Fig. [176]9H), without affecting its acetylation
   and phosphorylation (Supplementary Fig. [177]9I–K). Meanwhile, IFNγ
   stimulation significantly increased the acetylation of Munc13-4
   (Supplementary Fig. [178]9L), without influencing its phosphorylation
   and ubiquitylation (Supplementary Fig. [179]9M–O). These results
   indicate that both HRS ubiquitylation and Munc13-4 acetylation induced
   by IFNγ may cooperate to regulate PD-L1 sorting onto exosomes.
   To explore the mechanism that regulates Munc13-4 acetylation, we
   individually expressed six common acetyltransferases—GCN5, PCAF, CBP,
   P300, TIP60, and HBO1—with Munc13-4 in HEK293T cells and found that CBP
   and P300 are able to acetylate Munc13-4 (Supplementary Fig. [180]10A).
   Knockout of CBP—rather than P300—reduced the acetylation of endogenous
   Munc13-4 in SUM159 cells, both in the absence and presence of IFNγ
   (Supplementary Fig. [181]10B). In addition, IFNγ stimulation promoted
   the translocation of CBP from the nucleus to the cytoplasm
   (Supplementary Fig. [182]10C, D). These data indicate that CBP acts as
   a physiological acetyltransferase for Munc13-4. Then, we screened for
   the deacetylase mediating Munc13-4 deacetylation using deacetylase
   inhibitors. We found that trichostatin A (TSA), an inhibitor of histone
   deacetylases (HDACs)^[183]30, increased Munc13-4 acetylation, while
   nicotinamide (NIC), an inhibitor of class-III sirtuin deacetylases
   (SIRTs)^[184]30, had minimal effect (Supplementary Fig. [185]10E).
   Among the HDAC1–8 isoforms, expression of HDAC3 and HDAC4, but not the
   other HDACs, reduced Munc13-4 acetylation in HEK293T cells
   (Supplementary Fig. [186]10F–H). Notably, knockout of HDAC3, rather
   than HDAC4, increased the acetylation level of endogenous Munc13-4,
   regardless of IFNγ treatment (Supplementary Fig. [187]10I), suggesting
   that HDAC3 serves as a physiological deacetylase for Munc13-4. It is
   noteworthy that the deletion of CBP and HDAC also influenced Munc13-4
   expression, as evidenced by a reduction in Munc13-4 transcription
   (Supplementary Fig. [188]10J, L) and a corresponding decrease in the
   total amount of Munc13-4 (Supplementary Fig. [189]10K, M). We clarify
   that to ensure the accuracy of the data, our evaluation of the
   acetylation level of Munc13-4 in the experiments mentioned above was
   performed with the condition that the total amounts of Munc13-4 samples
   remained consistent (see “Methods”).
   We proceeded to investigate the impact of Munc13-4 acetylation on the
   sorting of PD-L1 onto exosomes. In the presence of IFNγ, the knockout
   of CBP and HDAC3 did not influence the total amount of PD-L1
   (Supplementary Fig. [190]10N, O) but displayed opposing effects on
   PD-L1 sorting. Specifically, the knockout of CBP, which reduces
   Munc13-4 acetylation, led to a significant increase in the abundance of
   PD-L1 on exosomes (Fig. [191]6A, B). In contrast, the knockout of
   HDAC3, which enhances Munc13-4 acetylation, resulted in a remarkable
   decrease in PD-L1 abundance on exosomes (Fig. [192]6C, D). Hence, these
   results consistently suggest that Munc13-4 acetylation inhibits the
   sorting of PD-L1 onto exosomes.
Fig. 6. IFNγ-induced modifications of Munc13-4 and HRS exert opposing effects
on PD-L1 sorting.
   [193]Fig. 6
   [194]Open in a new tab
   WB analysis on equal numbers of exosomes from control and CBP KO SUM159
   cells under IFNγ treatment (A) and quantification of blot band
   intensities (B) (n = 3 biological replicates). WB analysis on equal
   numbers of exosomes from control and HDAC3 KO SUM159 cells under IFNγ
   treatment (C) and quantification of blot band intensities (D) (n = 3
   biological replicates). WB analysis on equal numbers of exosomes from
   control and NEDD4L knockdown (KD) SUM159 cells under IFNγ treatment (E)
   and quantification of blot band intensities (F) (n = 3 biological
   replicates). WB analysis on equal numbers of exosomes from SUM159 cells
   co-treated with IFNγ and PR-619 or DMSO (G) and quantification of blot
   band intensities (H) (n = 3 biological replicates). I Representative
   confocal images of indicated SUM159 cells in the PLA (3 independent
   experiments). Scale bar, 10 μm. J Quantification of the mean number of
   PLA puncta per cell (n = 3 independent experiments). WB analysis on
   equal numbers of exosomes from indicated SUM159 cells (K) and
   quantification of blot band intensities (L) (n = 3 biological
   replicates). M Co-IP/IB analysis in HEK293T cells transfected with
   indicated constructs (n = 3 biological replicates). N IP/IB analysis in
   different SUM159 cells ± IFNγ treatment (n = 3 biological replicates).
   In (A, C, E, G, K), samples were run on the same gel; in (M, N), IP and
   Input samples were derived from the corresponding same experiment, but
   different gels for IP and Input samples were processed in parallel.
   Data are represented as means ± SEM (B, D, F, H, J, L), and p-values
   were calculated by two-way ANOVA with multiple comparisons (B, D, F, H,
   J, L). Source data are provided as a [195]Source data file.
   On the other hand, IFNγ stimulation induces HRS deubiquitylation, and
   we next investigate the effect of HRS ubiquitylation on PD-L1 sorting.
   To identify the E3 ligases responsible for HRS ubiquitylation, we
   utilized the UbiBrowser 2.0 database
   ([196]http://ubibrowser.bio-it.cn/ubibrowser_v3/)^[197]31. Among the
   identified and predicted E3 ligases, including NEDD4^[198]32,
   NEDD4L^[199]33, SH3RF1^[200]34, ITCH, CBL, and PRKN, we found that
   NEDD4L efficiently catalyzed HRS ubiquitylation in HEK293T cells
   (Supplementary Fig. [201]10P). In SUM159 cells, NEDD4L knockdown led to
   a significant reduction in HRS ubiquitylation, regardless of IFNγ
   stimulation (Supplementary Fig. [202]10Q). Notably, under IFNγ
   treatment, NEDD4L knockdown resulted in a marked increase in the amount
   of PD-L1 on exosomes (Fig. [203]6E, F) without altering total amount of
   PD-L1 (Supplementary Fig. [204]10Q). Furthermore, expression of several
   common deubiquitinases, including STAMBPL1, CYLD, USP11, USP7, and
   USP8, was found to reduce NEDD4L-mediated HRS ubiquitylation in HEK293T
   cells (Supplementary Fig. [205]10R), suggesting that the regulation of
   HRS deubiquitylation involves multiple deubiquitinases. Treatment with
   PR-619, a broad-spectrum reversible inhibitor of ubiquitin
   isopeptidases^[206]35,[207]36, induced a substantial increase in HRS
   ubiquitylation in SUM159 cells independent of IFNγ (Supplementary
   Fig. [208]10S). However, PR-619 treatment resulted in a significant
   reduction in the abundance of PD-L1 on exosomes under IFNγ stimulation
   (Fig. [209]6G, H), without influencing the overall PD-L1 level
   (Supplementary Fig. [210]10S). Collectively, these findings demonstrate
   that HRS deubiquitylation facilitates the sorting of PD-L1 onto
   exosomes.
IFNγ-induced modifications of Munc13-4 and HRS regulate PD-L1 binding
   The finding that Munc13-4 acetylation inhibits PD-L1 sorting suggests
   that it impairs the interaction between Munc13-4 and PD-L1. Indeed, the
   enhanced acetylation of Munc13-4, driven by the expression of CBP, led
   to a reduction in its interaction with PD-L1 (Supplementary
   Fig. [211]11A), as well as a weakened association of PD-L1 with HRS
   (Supplementary Fig. [212]11B), while not affecting the interaction
   between HRS and Munc13-4 (Supplementary Fig. [213]11B). This suggests
   that Munc13-4 serves as a central hub for the formation of the
   HRS–Munc13-4–PD-L1 complex. Interestingly, the mutant deleting the
   C-terminus of Munc13-4 (termed Munc13-ΔC, lacking residues 1049–1090)
   significantly diminished the acetylation by CBP (Supplementary
   Fig. [214]11C, D), indicating that the acetylation sites are located
   within this region. Furthermore, Munc13-ΔC failed to bind PD-L1
   (Supplementary Fig. [215]11E), suggesting that the residues within
   1049–1090 are crucial for both acetylation and PD-L1 binding. However,
   Munc13-ΔC retained its ability to bind HRS. Further screening revealed
   that residues 546–782 within the MUN domain of Munc13-4 mediate HRS
   interaction (Supplementary Fig. [216]11F).
   We aimed to identify the acetylation sites on Munc13-4. The region
   between residues 1049 and 1090 contains two lysine (K) residues, and
   mass spectrometry analysis revealed that both K1062 and K1079 were
   acetylated in the presence of CBP (Supplementary Fig. [217]11G, H).
   Single mutations of either K1062 or K1079 to arginine (K1062R or
   K1079R) preserved the positive charge but impaired acetylation.
   Moreover, the double mutant (K1062R/K1079R, termed KKRR) nearly
   completely abolished CBP-mediated acetylation of Munc13-4
   (Supplementary Fig. [218]11I), confirming our mass spectrometry
   findings. Detected by co-IP, mutating either K1062 or K1079 in Munc13-4
   to glutamine (Q), which mimics acetylation, did not affect its
   interaction with PD-L1 (Supplementary Fig. [219]11J). However, double
   mutations (K1062Q/K1079Q, termed KKQQ) significantly impaired this
   interaction (Supplementary Fig. [220]11J, K). In contrast, the KKRR
   mutant, which abolishes acetylation, did not affect the Munc13-4–PD-L1
   interaction (Supplementary Fig. [221]11K).
   In Munc13-4 knockout SUM159 cells, where the endogenous interaction
   between HRS and PD-L1 was severely damaged, expression of Munc13-4 WT
   rescued their endogenous interaction to the level comparable to that in
   control cells (Fig. [222]6I, J). In contrast, Munc13-4 KKQQ only
   slightly restored the HRS–PD-L1 interaction (Fig. [223]6I, J), while
   Munc13-4 KKRR substantially rescued this interaction (Fig. [224]6I, J).
   Similar results were observed in HEK293T cells (Supplementary
   Fig. [225]11L). Furthermore, the decrease in the PD-L1 abundance on
   exosomes observed upon Munc13-4 deletion was fully restored by the
   expression of either Munc13-4 WT or the KKRR mutant, but not by the
   KKQQ mutant (Fig. [226]6K, L). Overall, CBP-mediated acetylation of
   Munc13-4 at K1062 and K1079 disrupts its interaction with PD-L1, which
   in turn impairs the HRS–PD-L1 interaction and hinders the sorting of
   PD-L1 onto exosomes.
   We next investigated the mechanism by which HRS deubiquitylation
   promotes PD-L1 sorting onto exosomes. Expression of NEDD4L in HEK293T
   cells, which significantly enhanced HRS ubiquitylation, notably
   suppressed the interaction between HRS and PD-L1 (Fig. [227]6M). In
   contrast, additional expression of USP11, which decreased HRS
   ubiquitylation, restored the HRS–PD-L1 interaction (Fig. [228]6M).
   Notably, the ubiquitylation status of HRS did not affect its
   association with Munc13-4 (Supplementary Fig. [229]11M). These findings
   suggest that HRS ubiquitylation negatively regulates its interaction
   with PD-L1. In SUM159 cells, deletion of Munc13-4 also led to a
   reduction in HRS ubiquitylation, independent of IFNγ stimulation
   (Fig. [230]6N), which likely compensates for the diminished HRS–PD-L1
   interaction caused by Munc13-4 deficiency, suggesting a potential
   compensatory mechanism in PD-L1 sorting in the absence of Munc13-4.
A peptide that disrupts PD-L1–Munc13-4 interaction inhibits tumor growth
   Given the critical role of Munc13-4 in sorting PD-L1 to exosomes, we
   sought to disrupt their interaction in tumor cells to mitigate immune
   evasion. Targeting the PD-L1–Munc13-4 interaction might be a promising
   therapeutic strategy. Our co-IP analysis identified both the
   cytoplasmic motif and the transmembrane domain of PD-L1 as essential
   for binding to Munc13-4 (Supplementary Fig. [231]12A). To develop a
   peptide inhibitor targeting this interaction interface, we selected an
   18-residue segment (residues 256–273) spanning the
   transmembrane–cytoplasmic junction of PD-L1. This region encompasses
   key residues from both domains identified by co-IP and was chosen for
   its suitable length and favorable biochemical properties, such as
   predicted solubility.
   Next, we tested whether the 256–273 sequence could competitively
   inhibit the interaction between PD-L1 and Munc13-4. Overexpressing this
   sequence in HEK293T cells significantly disrupted their interaction
   (Supplementary Fig. [232]12B). This disruption also affected the
   interaction between HRS and PD-L1, but did not impact the HRS–Munc13-4
   interaction (Supplementary Fig. [233]12C). Further exploration utilized
   cell-penetrating peptide (CPP)^[234]37,[235]38 to deliver the PD-L1
   256–273 peptide (P-pep) and a scrambled version (S-pep) (Fig. [236]7A).
   Treatment of HEK293T cells with P-pep, compared to S-pep, significantly
   inhibited the ectopic interaction between PD-L1 and Munc13-4
   (Fig. [237]7B) and consequently disrupted the interaction between HRS
   and PD-L1 (Fig. [238]7C), while the HRS–Munc13-4 interaction remained
   unaffected (Fig. [239]7C). In addition, P-pep treatment of SUM159 cells
   led to a marked decrease in PD-L1 levels on exosomes (Fig. [240]7D),
   suggesting that competitive inhibition of PD-L1 sorting is a viable
   approach.
Fig. 7. A peptide disrupting PD-L1–Munc13-4 interaction inhibits tumor
growth.
   [241]Fig. 7
   [242]Open in a new tab
   A Diagram of the sequences for P-pep and S-pep. P-pep comprises a
   cell-penetrating peptide (CPP) fused to the human PD-L1 256–273 motif,
   whereas S-pep consists of a CPP linked to a scrambled sequence
   containing the same amino acid composition as the human PD-L1 256–273
   motif. B Co-IP/IB analysis in HEK293T cells transfected with indicated
   constructs and incubated with P-pep or S-pep to examine the effect of
   P-pep on Munc13-4–PD-L1 interaction (n = 3 biological replicates). C
   Co-IP/IB analysis in HEK293T cells transfected with indicated
   constructs and incubated with P-pep or S-pep to assess the effect of
   P-pep on the interactions of HRS with PD-L1 and Munc13-4 (n = 3
   biological replicates). D WB analysis on the same number of exosomes
   secreted from equal number of SUM159 cells treated with S-pep or P-pep
   (n = 3 biological replicates). E Schematic of experimental design for
   the assessment of in vivo anti-tumor efficacy of P-pep. F Tumor growth
   curves of orthotopic mouse models of breast cancer treated with P-pep
   or S-pep (n = 9 mice). Flow cytometric quantification of the percentage
   of CD45^+CD3^+CD4^+ (G) and CD45^+CD3^+CD8^+ (H) T cells among total
   cells in tumors (n = 5 mice). Representative contour plots depicting
   CD45^+CD3^+CD4^+ (I) and CD45^+CD3^+CD8^+ (K) T cell populations within
   tumors, showing the expression of granzyme B (n = 5 mice).
   Quantification of the percentage of granzyme B^+ cells among
   CD45^+CD3^+CD4^+ (J) and CD45^+CD3^+CD8^+ (L) T cells within tumors
   (n = 5 mice). In (B, C), IP and Input samples were derived from the
   corresponding same experiment, but different gels for IP and Input
   samples were processed in parallel; in (D), all samples were run on the
   same gel. Data are represented as means ± SEM (F), box plots show the
   median (center line), interquartile range (box), minima and maxima
   (whiskers), and all individual data points (dots) (G, H, J, L), and
   p-values were all calculated by unpaired two-tailed t-test. Source data
   are provided as a [243]Source data file.
   We also evaluated the in vivo effects of P-pep on tumor growth using
   orthotopic breast cancer mouse models (Fig. [244]7E and Supplementary
   Fig. [245]12D). Mice treated with P-pep exhibited a significant delay
   in tumor growth compared to those receiving S-pep (Fig. [246]7F and
   Supplementary Fig. [247]12E, F). Immunophenotyping revealed increased
   infiltration of both CD4^+ and CD8^+ T cells in the tumors of
   P-pep-treated mice (Fig. [248]7G, H), along with enhanced cytotoxicity
   of tumor-infiltrating CD8^+ T cells (Fig. [249]7K, L), while CD4^+ T
   cell cytotoxicity remained unchanged (Fig. [250]7I, J). Comprehensive
   assessments, including body weight, blood routine tests, biochemical
   analysis, and histological evaluations of major organs (Supplementary
   Fig. [251]12G–P), indicated no noticeable toxicity from either P-pep or
   S-pep. Collectively, these results demonstrate that P-pep effectively
   targets the PD-L1–Munc13-4 interaction, reducing tumor-induced
   immunosuppression and inhibiting tumor growth without systemic side
   effects.
Discussion
   Tumor-derived exosomes carry PD-L1, which engages with PD-1 receptors
   on T cells, influencing immune responses within the tumor
   microenvironment and in distant sites, resulting in a more extensive
   immunosuppressive environment^[252]5,[253]8,[254]39. Therapeutic
   strategies aimed at counteracting the immunosuppressive effects of
   exosomal PD-L1 require the development of molecules that can
   effectively suppress its extracellular secretion. Hence, it is
   essential to identify the regulatory factors governing the secretory
   pathway of exosomal PD-L1 and to elucidate the underlying mechanisms.
   In this study, we reveal a critical role for Munc13-4 in modulating the
   immunosuppressive effects of exosomal PD-L1 by influencing its sorting
   and secretion. Specifically, deleting Munc13-4 in breast tumor cells
   significantly reduces both the number of secreted exosomes and the
   abundance of PD-L1 on these exosomes, which systematically enhances T
   cell-mediated anti-tumor responses, suppresses tumor progression, and
   improves the efficacy of immune checkpoint inhibitors. The underlying
   mechanisms involve i) formation of the Munc13-4–PD-L1–HRS ternary
   complex, which promotes efficient sorting of PD-L1 to MVBs and loading
   onto exosomes; ii) assembly of the Munc13-4–Rab27a complex, which
   enables proper MVB docking; and iii) cooperation between Munc13-4 and
   the SNARE complex comprising syntaxin-4, SNAP-23 and VAMP-7, which
   facilitates MVB fusion with the plasma membrane to release
   PD-L1-containing exosomes. Notably, employing a specially designed
   peptide to disrupt the Munc13-4–PD-L1 interaction, thereby impairing
   PD-L1 sorting, significantly enhances anti-tumor immunity and slows
   tumor growth in vivo (Fig. [255]8). This underscores the potential of
   targeting the Munc13-4–PD-L1 axis as an effective approach to augment
   the efficacy of immune checkpoint inhibitors.
Fig. 8. Mechanistic model of Munc13-4-mediated tumor immune evasion through
the regulation of PD-L1 sorting and secretion via exosomes.
   [256]Fig. 8
   [257]Open in a new tab
   Schematic illustration showing that Munc13-4 collaborates with HRS,
   Rab27, and SNAREs to regulate PD-L1 sorting and secretion via exosomes.
   Loss of Munc13-4 in breast tumors enhances T cell-mediated anti-tumor
   immunity, suppresses tumor growth, and improves the efficacy of immune
   checkpoint inhibitors. Mechanistically, Munc13-4 regulates PD-L1
   sorting by forming a ternary complex with PD-L1 and HRS. IFNγ
   stimulation modifies Munc13-4 and HRS, establishing a dynamic
   regulatory mechanism that enables tumor cells to adapt to immune
   pressure by modulating exosomal PD-L1 sorting. Downstream of sorting,
   Munc13-4 engages Rab27a to regulate MVB tethering and promotes SNARE
   complex assembly, thereby facilitating MVB–plasma membrane fusion and
   exosome release. Therapeutically, a designed peptide that disrupts the
   Munc13-4–PD-L1 interaction impairs PD-L1 sorting, resulting in enhanced
   anti-tumor immunity and reduced tumor growth in vivo.
   Munc13-4 is ubiquitously expressed in various types of cells and
   contains a MUN domain flanked by two C[2] domains^[258]40. Munc13-4 has
   attracted significant attention for its role in the exocytosis of
   secretory granules in immune cells. Inherited variants of Munc13-4 have
   been associated with Familial Hemophagocytic Lymphohistiocytosis Type 3
   (FHL3), a rare autosomal recessive disorder characterized by impaired
   granule exocytosis^[259]16,[260]41. Munc13-4 is also involved in
   exosome secretion in tumor cells^[261]20. Building upon this existing
   knowledge, our work contributes important insights by revealing that
   Munc13-4 is upregulated in various tumor cells and regulates multiple
   steps toward MVB exocytosis, leading to exosome secretion, including
   MVB docking and fusion with the plasma membrane. This emphasizes its
   broader role in exocytosis across different cell types and highlights
   its potential as a biomarker for cancer diagnosis and prognosis. By
   regulating exosome secretion in tumor cells, Munc13-4 impacts the tumor
   microenvironment and facilitates communication between tumor and immune
   cells, influencing tumor growth, metastasis, and anti-tumor immunity.
   The accumulation of MVBs in Munc13-4 knockout cells suggests that
   Munc13-4 is crucial for the docking of MVBs with the plasma membrane.
   While previous studies have underscored the importance of the
   Munc13-4–Rab27a complex in regulating secretory granule
   docking^[262]19,[263]23,[264]24, the underlying mechanisms remain
   unclear due to the absence of structural information. In this study, we
   present the cryo-EM structure of the Munc13-4–Rab27a complex, revealing
   a previously unexplored binding interface essential for complex
   stability. Specifically, residues F46, W73, and F88 in Rab27a, along
   with residues N739/T740 and V660/K661 in Munc13-4, maintain this
   stability and mediate MVB docking with the plasma membrane. Notably,
   the Rab27a-binding surface on Munc13-4 is located within the middle of
   the MUN domain (spanning residues 651–778), rather than in an
   N-terminal sequence connecting the C[2]A and MUN domain as reported
   previously^[265]19,[266]24. Indeed, mutations in F46 and W73 in Rab27a
   and T740 in Munc13-4 have been previously linked to Griscelli Syndrome
   type 2 (GS2)^[267]42,[268]43 and FHL3^[269]44. Consistent with the
   structural data, introducing the mutations that disrupt the binding
   surface between Munc13-4 and Rab27a significantly increases the
   mobility of MVBs beneath the plasma membrane, which leads to impaired
   exosome secretion. Therefore, our results support the notion that
   Munc13-4 serves as the effector of Rab27a, stabilizing MVB docking
   close to the plasma membrane, thereby promoting exosome secretion.
   Together, Munc13-4–Rab27a-mediated MVB docking model may represent a
   common mechanism governing exosome secretion in various tumor cells.
   Our previous study has identified that the SNARE complex that mediates
   MVB fusion with the plasma membrane in tumor cells is composed of
   syntaxin-4, SNAP-23, and VAMP-7^[270]12. Here, we observed an
   accelerated effect of Munc13-4 on SNARE complex assembly and
   SNARE-mediated membrane fusion, which is dependent on its interaction
   with SNAP-23 (Q[bc]-SNARE) and VAMP-7 (R-SNARE). This suggests an
   important role for Munc13-4 in chaperoning the proper conformation of
   SNAREs and/or in stabilizing the SNARE complex. The SNARE-chaperoning
   role of Munc13s, particularly the well-studied isoform Munc13-1, has
   been extensively documented^[271]45–[272]49. While Munc13-4 shares
   similar domain structures with Munc13-1, their mechanisms in assisting
   SNARE complex assembly and membrane fusion may differ significantly.
   Syntaxin-4 is known to adopt a closed conformation similar to
   syntaxin-1^[273]50,[274]51. However, Munc13-4 lacks the hydrophobic
   core (the NF pocket) in its MUN domain^[275]46, making it unlikely to
   catalyze the opening of syntaxin-4. In this regard, the activation of
   syntaxin-4 may depend on the interaction between its N-peptide and the
   corresponding SM protein, Munc18-3^[276]52,[277]53. Our observation
   that Munc13-4 does not interact with syntaxin-4 further corroborates
   the notion. As a member of the CATCHR protein family, our finding that
   Munc13-4 binds to Q[bc]- and R-SNAREs is expected. For example,
   Munc13-1 binds to Q[bc]-SNARE SNAP-25^[278]54 and R-SNARE
   VAMP2^[279]48, aiding their assembly into the SNARE complex. The Dsl1
   complex interacts with Q[b]-SNARE Sec20 and Q[c]-SNARE Use1,
   stabilizing the complex’s conformation^[280]55. The GARP complex
   subunit Vps51 binds to Q[c]-SNARE Tlg1, likely promoting its connection
   to SNARE bundles^[281]56,[282]57. Similarly, the Exocyst complex
   subunit Sec6 interacts with Q[bc]-SNARE Sec9, facilitating both binary
   and ternary SNARE complex formation^[283]58,[284]59. Together with our
   findings, these evidences suggest that Munc13-4 promotes SNARE complex
   formation during MVB fusion through mechanisms shared by various CATCHR
   family members.
   The observation that Munc13-4 deletion does not alter the overall
   levels of PD-L1 on the plasma membrane but reduces the abundance of
   PD-L1 on secreted exosomes indicates a previously unrecognized role for
   Munc13-4 in cargo sorting within the endosomal system. HRS recognizes
   ubiquitinated proteins through its ubiquitin-interacting motif (UIM),
   which is essential for initiating cargo sorting during ESCRT-mediated
   MVB biogenesis^[285]60,[286]61. While HRS was reported to sort cargoes
   such as interleukin-2 receptor beta (IL-2Rβ)^[287]62 and PD-L1^[288]27
   independently of the UIM-ubiquitin interaction, the mechanism is not
   well understood. Our findings show that Munc13-4 independently binds
   both PD-L1 and HRS, but HRS cannot bind to PD-L1 without Munc13-4,
   highlighting Munc13-4’s critical role in recognizing PD-L1 before HRS
   interaction. This suggests that Munc13-4 mediates the recruitment of
   HRS to PD-L1, initiating a series of events that involve other ESCRT
   complexes and proteins required for the formation of PD-L1-containing
   ILVs. Therefore, the formation of a ternary complex consisting of
   Munc13-4, HRS, and Munc13-4-binding cargoes (e.g., PD-L1) may represent
   a ubiquitin-independent cargo sorting mechanism, where Munc13-4 and HRS
   work together to enable the proper sorting and packaging of cargoes
   into exosomes, facilitating its subsequent secretion. Further research
   is required to identify the full spectrum of exosomal cargo proteins
   sorted by Munc13-4. Overall, Munc13-4 works in conjunction with HRS,
   Rab27a, and SNAREs to establish an “assembly line” that effectively
   manages the processes of cargo sorting, packaging, trafficking, and
   release. This collaborative mechanism is critical for ensuring the
   efficient secretion of various proteins, including those involved in
   immune responses, highlighting the importance of Munc13-4 in exosomal
   biology.
   In the tumor microenvironment, tumor-infiltrating lymphocytes can
   secrete IFNγ to stimulate anti-tumor immune responses and induce tumor
   cell apoptosis^[289]28,[290]29. However, tumor cells can exploit IFNγ
   to reduce anti-tumor immunity by increasing PD-L1 levels on exosomes.
   Our findings reveal that IFNγ exerts dual and opposing effects on PD-L1
   sorting onto exosomes by modulating Munc13-4 acetylation and HRS
   deubiquitylation. IFNγ triggers the translocation of the
   acetyltransferase CBP from the nucleus to the cytoplasm, leading to
   acetylation of K1062/K1079 at the C-terminal end of Munc13-4. The
   acetylation of K1062/K1079 disrupts the Munc13-4–PD-L1 interaction,
   thereby reducing PD-L1 sorting. Notably, these two lysine residues are
   unique to Munc13-4 and not conserved among Munc13s, underscoring
   Munc13-4’s distinct functional role in cargo recognition. Meanwhile,
   IFNγ stimulates HRS deubiquitylation, which enhances the HRS–PD-L1
   interaction to increase PD-L1 sorting, likely due to a conformational
   change that leads to the activation of HRS^[291]63. These opposing
   effects indicate a nuanced regulatory mechanism where tumor cells may
   adapt to immune pressure by manipulating the sorting of PD-L1. Besides
   ubiquitylation, ERK-mediated phosphorylation of HRS also influences
   PD-L1 sorting^[292]27. Understanding these subtle regulatory mechanisms
   can help develop targeted therapies, such as modulating Munc13-4
   acetylation through CBP or HDAC3, or altering HRS ubiquitylation and/or
   phosphorylation via NEDD4L, deubiquitinases, or ERK. Recent studies
   suggest that exosomal PD-L1 levels in the blood of cancer patients
   could serve as a potential biomarker for predicting responses to immune
   checkpoint blockade (ICB) therapies^[293]5–[294]7. The variability in
   PD-L1 levels on circulating exosomes among different patients may be
   linked to differences in Munc13-4 acetylation and HRS
   ubiquitylation/phosphorylation. Our findings could provide valuable
   insights for guiding ICB treatment decisions by analyzing Munc13-4
   acetylation and HRS ubiquitylation in pathological tissue sections.
   Tumor-derived exosomes containing PD-L1 have systemic immunosuppressive
   effects^[295]8,[296]39, making the reduction of exosomal PD-L1
   secretion a promising therapeutic strategy. Genetic blockade of overall
   exosome secretion has proven effective in slowing tumor growth and
   enhancing immunotherapy outcomes. For instance, loss of Rab27a in tumor
   cells inhibits exosome secretion, decreasing the release of all
   exosomal cargoes, which leads to reduced tumor growth and improved T
   cell anti-tumor activity^[297]8,[298]64, similar to the effects
   observed with Munc13-4 in our study. However, while reducing overall
   exosome secretion can lessen their immunosuppressive effects, some
   evidence suggests that exosomes may also offer therapeutic benefits in
   immunotherapy. For example, exosomes can deliver tumor-associated
   antigens (TAAs) to dendritic cells, boosting T cell activation and
   promoting anti-tumor responses^[299]65. Exosomes carrying TAA-MHC
   complexes can directly stimulate antigen-specific T cell
   activation^[300]66. Therefore, selectively targeting PD-L1 sorting onto
   exosomes could eliminate their immunosuppressive effects while
   retaining their immunostimulatory potential. Our findings pinpoint the
   binding sites necessary for the Munc13-4–PD-L1 interaction that is
   specific for PD-L1 sorting. Disrupting this interaction with a designed
   peptide effectively reduces PD-L1 enrichment on exosomes without
   affecting overall exosome secretion. In vivo, this peptide treatment
   significantly enhances T cell function and inhibits tumor growth
   without major side effects. The peptide targets the PD-L1 motif that
   interacts with Munc13-4, allowing for selective PD-L1 targeting and
   minimizing the risk of immune dysfunction from non-specific peptide
   uptake by immune cells. Overall, this peptide holds great promise as a
   therapeutic agent for modulating immune responses.
   In conclusion, we elucidate the Munc13-4-dependent mechanisms that
   govern the secretion of PD-L1 via exosomes in breast cancer cells,
   highlighting the functional complexity of Munc13-4 and enhancing our
   understanding of exosome biogenesis and secretion. Our in vivo findings
   offer valuable insights into the regulation of PD-L1 secretion,
   suggesting promising therapeutic strategies to improve patient outcomes
   in cancer treatment.
Methods
Ethics statement
   For mice: BALB/c, BALB/Nude, and NOD/SCID mice (female, 5–7-week-old)
   were purchased from Beijing Vital River Laboratory Animal Technology
   Co., Ltd. (Beijing, China). The mice were housed in a controlled animal
   facility under consistent environmental conditions, including a room
   temperature of 22  ±  1 °C, relative humidity of 40–70%, and a 12-h
   light/dark cycle. Food and water were provided ad libitum. Mice were
   randomly assigned at the start of each experiment. All experimental
   procedures were conducted in compliance with the guidelines and with
   approval from the Institutional Animal Care and Use Committee (IACUC)
   of Tongji Medical College, Huazhong University of Science and
   Technology (Wuhan, China).
   For patients’ samples: formalin-fixed, paraffin-embedded human tissue
   arrays (HOrgC180PG01-2 and HBreD180Bc01-2) were obtained from Shanghai
   Outdo Biotech Co., Ltd. (China). The HOrgC180PG01-2 array comprises 180
   tissue cores derived from 91 patients across 14 tumor types, and the
   HBreD180Bc01-2 array includes 180 tissue cores from 150 patients with
   triple-negative breast cancer. Both tissue arrays were subjected to
   immunohistochemical (IHC) analysis, with detailed clinical information
   retrieved from the company’s website
   ([301]https://www.superchip.com.cn/). All the human tissue samples were
   collected with informed consent from the donors, and their use was
   approved by the Ethics Committee of Shanghai Outdo Biotechnology Co.,
   Ltd., in accordance with relevant ethical guidelines and regulations.
Bacterial strains
   Escherichia coli strains BL21(DE3) (Thermo), DH10Bac (Gibco), and DH5α
   (Thermo) were cultured in Luria-Bertani (LB) broth at 37 °C with
   shaking at 200 rpm. The media was supplemented with appropriate
   antibiotics: ampicillin (100 µg/ml), kanamycin (50 µg/ml), gentamicin
   (7 µg/ml), or tetracycline (10 µg/ml), as necessary.
Cell culture
   HEK293T and 4T1 cells were obtained from American Type Culture
   Collection (ATCC). SUM159 were obtained from Pcocell. HEK293T and
   SUM159 cells were cultured in DMEM medium (Gibco, 11965092) with 10%
   FBS (Gibco, A5670701) and 1% penicillin-streptomycin solution
   (Proteintech, PR40022). 4T1 cells were cultured in RPMI 1640 medium
   (Gibco, 11875093) supplemented with 10% FBS (Gibco, A5670701) and 1%
   penicillin-streptomycin (Proteintech, PR40022). These cells were
   cultured in a humidified incubator (Thermo) at 37 °C with 5% CO2. Sf9
   cells (Gibco) were cultured in SIM-SF Expression Medium
   (SinoBiological, MSF1) at 27 °C, 125 rpm.
Online data acquisition and analysis
   The differential expression of Munc13-4 between tumor and normal
   tissues was analyzed utilizing TIMER2.0
   ([302]http://timer.cistrome.org/)^[303]21,[304]22. The reported and
   predicted ubiquitin ligase (E3) of HRS was searched using UbiBrowser
   2.0 ([305]http://ubibrowser.bio-it.cn/ubibrowser_v3/)^[306]31.
Plasmids
   The coding sequences for full-length human CD63, PD-L1, and Munc13-4
   were inserted into the NPY-td-Orange2 vector. Full-length human
   Munc13-4, PD-L1, and various mutants of Munc13-4, including
   K1062R/K1079R, K1062Q/K1079Q, V660A/K661A, and N739G/T740G, along with
   truncated forms of Munc13-4 (residues 1–1048, 1–910, 1–782, 1–546,
   1–287, and 1–108), were cloned into the pEGFP-N3 vector (Clontech).
   Similarly, full-length human HRS, STAM, and Rab27a, along with Rab27a
   mutants (F46S, W73S, and F88S), were cloned into the pEGFP-C1 vector
   (Clontech).
   Constructs for full-length human Munc13-4, PD-L1, HRS, and mutant
   Munc13-4 (K1062R/K1079R, K1062Q/K1079Q), as well as truncated PD-L1
   (residues 256–273), histone acetyltransferase domains of human GCN5
   (residues 503–656), PCAF (residues 503–651), CBP (residues 1323–1700),
   P300 (residues 127–1663), TIP60 (residues 227–504), and HBO1 (residues
   332–607), and full-length human NEDD4, ITCH, CBL, SH3RF1, PRKN, NEDD4L,
   STAMBPL1, CYLD, USP11, USP7, USP8, USP36, were generated in the
   pcDNA3.1- vector (Invitrogen) with an N-terminal Flag-tag. Histone
   deacetylase domains of human HDAC1 (residues 9–321), HDAC2 (residues
   9–322), HDAC3 (residues 3–316), HDAC4 (residues 655–1084), HDAC5
   (residues 684–1028), HDAC6 (residues 87–404 and 482–800), HDAC7
   (residues 518–865), and HDAC8 (residues 14–324) were also cloned into
   the pcDNA3.1- vector with an N-terminal HA-tag. Coding sequence of
   full-length human NEDD4L was cloned into the pcDNA3.1- vector
   (Invitrogen) with an N-terminal Strep-tag.
   Full-length human Munc13-4, Rab27a, and their respective mutants,
   including Munc13-4 (K1062R/K1079R, K1062Q/K1079Q, V660A/K661A,
   N739G/T740G) and Rab27a (F46S, W73S, F88S), were cloned into the
   pLV-EF1α-IRES-Hygro vector (Addgene). Full-length Munc13-4 and its
   mutants (V660A/K661A, N739G/T740G) were subcloned into the
   pFastBac^TMHT B vector (Invitrogen).
   Coding sequences for full-length human Rab27a, syntaxin-4, truncated
   syntaxin-4 (residues 1–275, Syx-4 ΔTM), SNAP-23 mutant (All cysteine
   residues were mutated to serine, SN-23-6CS), SN-23-6CS (S161C), VAMP-7
   SNARE motif (residues 123–187, A131C), VAMP-7 SNARE-TM motif (residues
   123–208), and PD-L1 (residues 19–290) were cloned into the pET-28a
   vector (Novagen). Finally, full-length Rab27a, Rab5, Rab7, and their
   mutants (Rab27a (F46S, W73S, F88S)), Syx-4 ΔTM (1–275), SN-23-6CS, and
   VAMP-7 SNARE motif (123–187) were subcloned into the pGEX-6P-1 vector
   (Cytiva).
Transfection
   Recombinant bacmid was transfected into Sf9 cells using the
   X-tremeGENE™ 9 DNA Transfection Reagent (Roche). For all other plasmid
   transfections, Hieff Trans® Liposomal Transfection Reagent (Yeasen,
   40802ES08) was employed, following the manufacturer’s instructions.
   Cells were prepared for subsequent experiments 24–36 h
   post-transfection.
Cell viability assay
   Cell proliferation of control and Munc13-4 knockout 4T1 cells was
   assessed using a CCK-8 kit (Vazyme, A311-01) according to the
   manufacturer’s instructions. Briefly, the same number of control and
   Munc13-4 knockout cells were plated onto 96-well culture plates, and
   cell viability was measured at 0, 24, and 48 h using the CCK-8 kit.
Histological analyses
   Tissue arrays were deparaffinized with heat at 60 °C for 30 min
   followed by two 15-min washes with xylene. Then, the paraffin sections
   were rehydrated by washing for 5 min in absolute ethanol I, absolute
   ethanol II, 85% alcohol, 75% alcohol, and distilled water in sequence.
   Following the procedures outlined in the “PTLink Quick Operation Guide”
   (Dako), slides were subjected to antigen retrieval using the specified
   instrument. Upon completion, slides were immersed in distilled water at
   room temperature for natural cooling for a minimum of 10 min.
   Subsequently, the slides were rinsed with PBST buffer. The diluted
   Munc13-4 primary antibody working solution (1:50, Santa Cruz,
   sc-271300) was applied, and the slides were incubated overnight at
   4 °C. The next day, the slides were removed from refrigeration and
   allowed to equilibrate to room temperature for 45 min before being
   washed with PBST buffer. Automated staining, including blocking,
   secondary antibody binding, and DAB color development, was performed
   using the DAKO automated immunohistochemistry staining system according
   to the “Autostainer Link 48 User Guide.” Counterstaining was conducted
   using hematoxylin for 1 min, followed by immersion in 0.25%
   hydrochloric acid alcohol (prepared with 400 ml of 70% ethanol and 1 ml
   of concentrated hydrochloric acid) for no less than 2 s. The slides
   were rinsed under running water for a minimum of 2 min, air-dried at
   room temperature, and mounted using neutral resin. Digitization of the
   slides was performed at ×20 magnification using the Aperio XT Scanner
   (Leica). To evaluate the protein expression levels of Munc13-4 in
   breast cancer tissues and adjacent normal tissues, immunohistochemistry
   (IHC) images were analyzed using IHC Profiler, an open-source plugin
   for ImageJ. The staining intensity for Munc13-4 was scored using a
   four-tier scoring system: 0 (negative), 1 (low positive), 2 (positive),
   and 3 (high positive).
   Samples of mouse heart, liver, spleen, lung, and kidney were fixed
   overnight in 4% formalin, embedded in paraffin, and cut into 4 mm
   consecutive sections. The paraffin sections were sequentially immersed
   in Environmental-Friendly Dewaxing Transparent Liquids I and II
   (Servicebio, G1128) for 20 min each, followed by treatment with
   anhydrous ethanol I and II for 5 min each. Subsequently, the sections
   were immersed in 75% ethanol for 5 min and thoroughly rinsed with tap
   water. Hematoxylin and eosin (H&E) staining was performed using the
   Hematoxylin-Eosin (H&E) HD Constant Dye Kit (Servicebio, G1076)
   according to the manufacturer’s instructions. The sections were then
   dehydrated through a graded series of absolute ethanol solutions (I,
   II, and III) for 2 min each, followed by sequential immersion in normal
   butanol I and II for 2 min each and clearing in xylene I and II for
   2 min each. Finally, the sections were sealed with neutral gum and
   scanned using the NanoZoomer S360 Digital Slide Scanner (Hamamatsu).
   For tissue immunofluorescence assays, paraffin-embedded tumor tissues
   from orthotopic mouse models of breast cancer sections were
   deparaffinized and rehydrated through a graded ethanol series, followed
   by washing in distilled water. Antigen retrieval was performed using
   EDTA Antigen Retrieval Solution (Beyotime, P0085) under high
   temperature and pressure conditions, and the sections were allowed to
   cool to room temperature before washing in Tris-buffered saline with
   0.05% Tween-20, pH 7.4 (TBST). Endogenous peroxidase activity was
   blocked using 3% H2O2, followed by washing in distilled water. The
   sections were then encircled with a hydrophobic pen and incubated with
   10% goat serum (Boster, AR1009) at 37 °C for blocking. For the first
   staining, a CD4 primary antibody (Abcam, RM1013, 1:50) diluted in TBST
   was applied, and the sections were incubated overnight at 4 °C. After
   washing, a secondary antibody, Goat Anti-Rabbit IgG H&L (HRP) (Abcam,
   ab205718, 1:4000) was added and incubated at 37 °C. Tyramide signal
   amplification (TSA) staining was performed using iFluor^® 488 tyramide
   working solution (AAT Bioquest, 45100), followed by washing in TBST.
   The slides underwent a second round of antigen retrieval in Improved
   Citrate Antigen Retrieval Solution (Beyotime, P0083) using a microwave,
   were cooled to room temperature, and washed again. Blocking was
   repeated with 10% goat serum at 37 °C. For the second staining, a CD8
   primary antibody (Abcam, RM1129, 1:100) diluted in TBST was applied,
   and the sections were incubated overnight at 4 °C. Following washing,
   the HRP-conjugated secondary antibody was added and incubated, and TSA
   staining was conducted using Cy3 tyramide working solution (AAT
   Bioquest, 11065). Nuclear staining was performed with DAPI (Solarbio,
   C0060) in the dark, followed by washing in TBST. Finally, the sections
   were mounted with Fluoromount-G^® (SouthernBiotech, 0100-01) and stored
   at 4 °C in the dark. The sections were imaged using Pannoramic SCAN II
   (3D HISTECH).
iTRAQ-based quantitative proteomics
   4T1 cells (control and Munc13-4 knockout) were plated in triplicate for
   proteomics. To extract proteins, an appropriate volume of SDS-free L3
   buffer supplemented with final concentration of 1 × Cocktail (EDTA
   contained) was added to the sample. The mixture was incubated on ice
   for 5 min, followed by the addition of DTT to achieve a final
   concentration of 10 mM. Ultrasonic disruption was performed to lyse the
   sample, and the lysate was centrifuged at 25,000 × g and 4 °C for
   15 min to remove insoluble debris. The supernatant was collected and
   further treated with DTT (final concentration of 10 mM), followed by
   incubation in a water bath at 56 °C for 1 h. Subsequently,
   iodoacetamide was added to the solution to a final concentration of
   55 mM, and the mixture was incubated in the dark for 45 min. A second
   centrifugation at 25,000 × g and 4 °C for 15 min was performed, and the
   supernatant, containing the extracted protein solution, was collected
   for downstream analyses.
   Protein samples (100 µg each) were digested with Trypsin Gold (Promega,
   V5280) at a protein-to-trypsin ratio of 20:1 (w/w) at 37 °C for 16 h.
   The resulting peptides were dried via vacuum centrifugation and
   reconstituted in 0.5 M TEAB. iTRAQ labeling was conducted following the
   manufacturer’s protocol for the 4-plex iTRAQ reagent kit
   (Sigma-Aldrich, 4374321). The labeled samples were combined in equal
   proportions and fractionated using high-performance liquid
   chromatography (HPLC) on a Thermo DIONEX Ultimate 3000 BioRS system
   equipped with a Durashell C18 column (5 µm, 100 Å, 4.6 × 250 mm, Welch
   Materials). A total of 20 fractions were collected for further
   analysis.
   Peptides separated by liquid chromatography were ionized using a
   nanoESI source and analyzed on a Q-Exactive HF X mass spectrometer
   (Thermo Fisher Scientific) operating in data-dependent acquisition
   (DDA) mode. Instrument parameters were configured as follows: the ion
   source voltage was set to 1.9 kV; the MS1 scan range was 350–1,500 m/z
   with a resolution of 60,000; and the MS2 scan range started at a fixed
   m/z of 100 with a resolution of 15,000. Precursor ion selection
   criteria included charge states between 2+ and 6+ and the top 20 most
   intense ions with signal intensities exceeding 10,000. Fragmentation
   was performed using higher-energy collisional dissociation (HCD), and
   the resulting fragments were detected in the Orbitrap analyzer. The
   dynamic exclusion duration was set to 30 s, and the automatic gain
   control (AGC) targets were 3 × 10^6 for MS1 and 1 × 10^5 for MS2.
   Protein identification was conducted utilizing the Mascot search engine
   (version 2.3.02; Matrix Science) against the Mus musculus subset of the
   NCBI non-redundant (NR) sequence databases. The search parameters were
   configured as follows: monoisotopic mass, peptide mass tolerance of
   20 ppm, fragment mass tolerance of 0.05 Da, trypsin as the digestion
   enzyme, allowance for one missed cleavage, and charge states of +2 and
   +3 for peptides. Variable modifications included Gln-> pyro-Glu
   (N-terminal Q), oxidation (M), and deamidation (NQ), while fixed
   modifications comprised carbamidomethylation (C) and iTRAQ8plex
   labeling (N-terminal and K). Protein quantification was performed using
   the automated software IQuant. Peptides with a confidence interval of
   95% were filtered based on a 1% false discovery rate (FDR), and
   confident proteins were required to include at least one unique
   peptide. Quantitative protein ratios were weighted and normalized using
   the median ratio in Mascot. Differentially expressed proteins (DEPs)
   between the control and Munc13-4 knockout groups were identified via
   t-tests, with results subjected to a 5% FDR correction. Proteins with
   expression fold changes ≥1.2 or ≤0.83 were classified as DEPs.
   Additionally, KEGG pathway enrichment analysis was performed for DEPs,
   and a heatmap was generated using an online platform for data analysis
   and visualization ([307]https://www.bioinformatics.com.cn/)^[308]67.
Cell treatment with IFNγ, PR-619, and peptide
   To explore the effects of IFNγ on the expression and exosomal sorting
   of PD-L1, 100 ng/ml Recombinant Human IFN-gamma Protein (Abclonal,
   RP01038) was added to SUM159 cells, and 100 ng/ml Recombinant Mouse
   IFN-gamma Protein (Abclonal, RP01070) was added to 4T1 cells,
   incubating for 24 h. To examine the effect of HRS ubiquitylation on the
   sorting of PD-L1 onto exosomes under IFNγ stimulation, SUM159 cells
   were co-treated with 100 ng/ml Recombinant Human IFN-gamma Protein
   (Abclonal, RP01038) and 8 μM PR-619 (MCE, HY-13814) for 24 h. To
   investigate the effects of the designed peptide on protein interactions
   and exosomal PD-L1 sorting, HEK293T and SUM159 cells were incubated
   with 10 μg/ml P-pep (Homo sapiens) or S-pep.
Western blot
   Cells or exosomes were lysed on ice in RIPA buffer (50 mM Tris-HCl, pH
   7.5, 150 mM NaCl, 1% Triton X-100) supplemented with a protease
   inhibitor cocktail (Topscience, C0001). Following a 20-min incubation,
   the lysates were centrifuged at 4 °C, 12,000 × g for 10 min. The total
   protein concentration in the supernatant was determined using a BCA
   Protein Quantification Kit (Yeasen, 20201ES76) to ensure consistent
   loading of different samples. Protein samples were then denatured by
   heating in diluted 1× SDS-PAGE Sample Loading Buffer (Yeasen,
   20315ES05) for 10 min at 100 °C. Proteins were separated by SDS-PAGE
   and subsequently transferred to PVDF membranes (Millipore, ISEQ00010).
   The membranes were blocked with 5% non-fat bovine milk in Tris-Cl
   buffer (150 mM NaCl, pH 7.2) containing 0.1% Tween-20, followed by
   incubation with the indicated primary antibody and a subsequent
   incubation with an HRP-conjugated secondary antibody. Immunodetection
   was carried out using the Super Sensitive ECL Luminescence Reagent
   (Meilunbio, MA0186-2). The integrated density of the blot bands was
   quantified and analyzed using ImageJ and Prism 6.0 software to assess
   the relative protein levels. The primary antibodies used in western
   blot assays are as follows: Munc13-4 Antibody (C-2) (Santa Cruz,
   sc-271300, 1:1000, for samples of human origin), UNC13D Monoclonal
   antibody (Proteintech, 67193-1-Ig, 1:5000, for samples of mouse
   origin), PD-L1/CD274 Rabbit mAb (Abclonal, A19135, 1:2000, for samples
   of human origin), Anti-PD-L1 antibody (abcam, ab213480, 1:2000, for
   samples of mouse origin), HGS Polyclonal antibody (Proteintech,
   10390-1-AP, 1:10000, for samples of both human and mouse origin),
   β-Actin Rabbit mAb (High Dilution) (Abclonal, AC026, 1:100000, for
   samples of both human and mouse origin), Alix Monoclonal antibody
   (Proteintech, 67715-1-Ig, 1:5000, for samples of both human and mouse
   origin), CD63 Antibody (MX-49.129.5) (Santa Cruz, sc-5275, 1:1000, for
   samples of both human and mouse origin), CD81 Antibody (B-11) (Santa
   Cruz, sc-166029, 1:1000, for samples of both human and mouse origin),
   Mouse anti-GFP-Tag mAb (Abclonal, AE012, 1:10000, species independent),
   Rabbit anti-GFP-Tag pAb (Abclonal, AE011, 1:10000, species
   independent), DYKDDDDK tag Polyclonal antibody (Proteintech,
   20543-1-AP, 1:10000, species independent), StrepII Tag Mouse Monoclonal
   Antibody (Beyotim, AF2924, 1:1000, species independent), Mouse anti
   HA-Tag mAb (Abclonal, AE008, 1:5000, species independent), Pan
   Acetylation Monoclonal antibody (Proteintech, 66289-1-Ig, 1:1000,
   species independent), Ubiquitin Antibody (P4D1) (Santa Cruz, sc-8017,
   1:1000, for samples of both human and mouse origin), pan
   Phospho-Serine/Threonine Rabbit Polyclonal Antibody (Beyotim, AF5725,
   1:1000, species independent), Pan Phospho-Tyrosine Mouse mAb (Abclonal,
   AP0973, 1:1000, for samples of both human and mouse origin),
   CBP/KAT3A/CREBBP Antibody (C-1) (Santa Cruz, sc-7300, 1:1000, for
   samples of both human and mouse origin), P300 Antibody (F-4) (Santa
   Cruz, sc-48343, 1:1000, for samples of both human and mouse origin),
   Histone Deacetylase 3 (HDAC3) Antibody (A-3) (Santa Cruz, sc-376957,
   1:1000, for samples of both human and mouse origin), Histone
   Deacetylase 4 (HDAC4) Antibody (B-5) (Santa Cruz, sc-365093, 1:1000,
   for samples of both human and mouse origin), NEDD4L Polyclonal antibody
   (Proteintech, 13690-1-AP, 1:3000, for samples of both human and mouse
   origin). The secondary antibodies used in western blot experiments are
   as follows: HRP-conjugated Goat anti-Rabbit IgG (H+L) (Abclonal, AS014,
   1:10000), HRP-conjugated Goat anti-Mouse IgG (H+L) (Abclonal, AS003,
   1:10000). The integrated density of blot strips was analyzed by ImageJ
   software to characterize the relative protein level.
Generation of gene-edited cell lines
   To generate gene knockout cell lines, the CRISPR-Cas9 system was
   employed. HEK293T cells were transfected with the lentiCRISPR v2
   plasmid, which contained a single-guide RNA (sgRNA) targeting the gene
   of interest, along with the psPAX2 and pMD2.G plasmids to produce
   lentivirus. After 36–48 h of transfection, the culture medium of the
   HEK293T cells was collected, centrifuged to remove cell debris, and the
   supernatant containing the lentivirus-sgRNA was used to infect SUM159
   or 4T1 cells. Following 48 h of infection, cells were selected with
   puromycin (MCE, HY-B1743) at a concentration of 2.0 μg/mL for 5–7 days.
   Limiting dilution was then performed to isolate single-cell clones from
   the infected SUM159 or 4T1 cells. Cells infected with lentivirus
   produced by HEK293T cells co-transfected with an empty lentiCRISPR v2
   vector, psPAX2, and pMD2.G plasmids served as controls. Successful
   knockout of the target SNARE protein was verified by western blot
   analysis.
   The sgRNAs used in this study are as follows:
   Mus-Munc13-4: GTGGCCTTCAGGCAAAATAC
   Hs-Munc13-4: TGAAGGTCTCGTCCCAGACG
   Hs-Rab27a: CCAAAGCTAAAAACTTGATG
   Hs-HRS: CTGCCTGCAGAGACAAGTGG
   Hs-P300: GTTCAATTGGAGCAGGCCGA
   Hs-CBP: CGCGTGACCAGTCATTTGCG
   Hs-HDAC3: GGTGAAGCCTTGCATATTGG
   Hs-HDAC4: GGAGCCCATTGAGAGCGATG
   Hs-NEDD4L: GGAGCCCATTGAGAGCGATG
   For the generation of gene-complemented cell lines, the
   pLV-EF1a-IRES-Hygro plasmid containing the full-length sequence of the
   gene of interest was utilized. HEK293T cells were transfected with the
   pLV-EF1a-IRES-Hygro plasmid along with the psPAX2 and pMD2.G plasmids
   to produce lentivirus. At 36–48 h post-transfection, the culture medium
   was collected, centrifuged to remove cell debris, and the resulting
   lentivirus-containing supernatant was used to infect the gene knockout
   SUM159 cells. Infected cells were selected using Hygromycin B (Sangon
   Biotech, A100607) at a final concentration of 500 μg/ml for 5–7 days.
   Single-cell clones were subsequently isolated through limiting
   dilution.
Immunoprecipitation (IP) and co-IP
   To investigate the post-transcriptional modifications of Munc13-4 and
   HRS, IP assays were conducted. Equal numbers of SUM159 cells were
   seeded onto 15-cm culture dishes, with one plate treated with 100 ng/ml
   IFNγ (ABclonal, RP01038) for 24 h. Following the incubation, cells were
   lysed on ice using RIPA buffer (50 mM Tris-HCl, pH 7.5, 150 mM NaCl, 1%
   Triton X-100) supplemented with a protease inhibitor cocktail
   (Topscience, C0001). The lysates were then centrifuged at 4 °C for
   10 min at 12,000 × g. The supernatant was incubated overnight at 4 °C
   with Munc13-4 Antibody (C-12) (Santa Cruz, sc-271301, 1:100) or HGS
   Polyclonal antibody (Proteintech, 10390-1-AP, 1:500), in conjunction
   with Protein A/G magnetic beads (Biolinkedin, L-1004), on a rotator. To
   identify the acetyltransferase of Munc13-4, GFP-tagged Munc13-4 was
   co-expressed with Flag-labeled GCN5 503–656, PCAF 503–651, CBP
   1323–1700, P300 127–1663, TIP60 227–504, or HBO1 332–607 in HEK293T
   cells. To investigate the type of deacetylase, HEK293T cells
   co-expressing GFP-Munc13-4 and Flag-CBP 1323–1700 were treated with
   0.2% DMSO, 5 mM nicotinamide (NIA) (MCE, HY-B0150) or 1 μM trichostatin
   A (TSA) (MCE, HY-15144) for 24 h. To identify the deacetylase of
   Munc13-4, GFP- Munc13-4 and Flag- CBP 1323–1700 was co-expressed with
   HA-fused HDAC1 9–321, HDAC2 9–322, HDAC3 3–316, HDAC4 655–104, HDAC5
   684–1028, HDAC6 87–404, HDAC6 482–800, HDAC7 518–865 or HDAC8 14–324 in
   HEK293T cells. To identify the E3 ligase of HRS, GFP-HRS was
   co-expressed with Flag-tagged NEDD4, ITCH, CBL, SH3RF1, PRKN, or NEDD4L
   in HEK293T cells. To identify the deubiquitinase of HRS, GFP-HRS and
   strep-NEDD4L were co-expressed with Flag-fused STAMBPL1, CYLD, USP11,
   USP7, USP8, or USP36 in HEK293T cells. The transfection of plasmids
   mentioned above was performed using Hieff Trans^® Liposomal
   Transfection Reagent (Yeasen, 40802ES03) according to the
   manufacturer’s protocol.GFP-Munc13-4 or GFP-HRS was immunoprecipitated
   from cell lysates using anti-GFP magnetic beads (Biolinkedin, L-1016)
   according to the manufacturer’s guidance. After incubation, the beads
   were washed three times with PBS. Subsequently, 1× SDS-PAGE Sample
   Loading Buffer (Yeasen, 20315ES05) was added, and the samples were
   heated at 100 °C for 10 min. The protein samples were then analyzed by
   western blotting.
   The deletion of either CBP or HDAC3 in SUM159 cells resulted in a
   significant decrease in Munc13-4 expression. To examine the effects of
   CBP and HDAC3 knockout on Munc13-4 acetylation, Munc13-4 was enriched
   from a cell population three times larger in the knockout groups
   compared to the control cells. The amount of immunoprecipitated
   Munc13-4 used for acetylation analysis was standardized based on the
   results of preliminary western blot analysis.
   To investigate protein-protein interactions, co-IP assays were
   conducted. Recombinant plasmids were transfected into SUM159 or HEK293T
   cells using Hieff Trans^® Liposomal Transfection Reagent (Yeasen,
   40802ES03). GFP-fused proteins were then enriched from cell lysates
   using anti-GFP magnetic beads (Biolinkedin, L-1016). Following
   enrichment, the beads were washed three times with PBS and subsequently
   heated in 1× SDS-PAGE Sample Loading Buffer (Yeasen, 20315ES05) at
   100 °C for 10 min. Next, the samples were analyzed by western blotting.
Isolation of EVs/exosomes
   For extracellular vesicle (EV)/exosome preparation, equal numbers of
   cells with the indicated genotypes were seeded onto 15-cm dishes. Once
   the cells adhered, they were rinsed with PBS and cultured for 24 h in
   DMEM supplemented with 10% exosome-depleted FBS (prepared by
   ultracentrifugation at 100,000 × g overnight to remove bovine
   vesicles). Conditioned medium was collected and subjected to stepwise
   centrifugation to eliminate contaminants. Specifically, the medium was
   centrifuged at 300 × g for 10 min at 4 °C to remove intact cells,
   followed by 2000 × g for 20 min to eliminate cellular debris, and then
   10,000 × g for 30 min to deplete larger vesicles. The clarified
   supernatant was ultracentrifuged at 100,000 × g for 70 min (Beckman
   Type 70Ti rotor) to pellet exosomes. The pellet was washed once with
   cold PBS and centrifuged again at 100,000 × g for 70 min to enhance
   purity. The final exosome pellet was resuspended in PBS or RIPA buffer
   for downstream assays.
Optiprep^TM density gradient centrifugation
   Pellets of EVs, obtained through ultracentrifugation from cell culture
   supernatants, were washed and resuspended in 200 µL of buffer
   containing 0.25 M sucrose, 10 mM Tris-Cl, and 1 mM EDTA (pH 7.4). The
   suspension was then transferred to an SW55Ti rotor tube (Beckman,
   344090), mixed in a 1:1 ratio with a 60% (wt/vol) Optiprep™ stock
   solution, and sequentially layered with 160 µL of a 20% (wt/vol)
   Optiprep™ solution and 150 µL of a 10% (wt/vol) Optiprep™ solution.
   Tubes were centrifuged for 1 h at 4 °C, 350,000 × g in an SW55Ti rotor
   (stopping without break). Following centrifugation, six 100 µL
   fractions were collected from the top of the gradient. These fractions
   were diluted with 600 µL of PBS and subjected to a second round of
   centrifugation for 1 h at 4 °C, 100,000 × g. The resulting pellets from
   the concentrated fractions were resuspended in 20 µL of PBS and
   analyzed by western blotting.
Transmission electron microscopy (TEM)
   The morphology of EVs was characterized by TEM. 20 μL of EVs suspension
   was carefully deposited onto a copper grid (EMCN, BZ11262a) and
   incubated for 3–5 min. Excess liquid was then removed with filter
   paper. Subsequently, 2% uranyl acetate was applied to the copper grid
   for 2–3 min, after which the excess solution was absorbed using filter
   paper, and the sample was allowed to air-dry at room temperature. The
   samples were then observed with TEM (HITACHI, HT7800). The number of
   ILVs and MVBs and the percentage of MVB-lysosome hybrids among total
   MVBs were scored manually, and the diameter of MVBs was measured by
   ImageJ software.
Nanoparticle tracking analysis (NTA)
   Exosome size distribution and concentration were assessed using a
   NanoSight NS300 system (Malvern) equipped with a high-sensitivity sCMOS
   camera. Samples were recorded in triplicate, with three 60-s videos
   collected for each preparation. The captured data were analyzed using
   NanoSight NTA software, which determines particle size and
   concentration based on tracking Brownian motion and calculating the
   corresponding diffusion coefficients.
Tunable resistive pulse sensing (TRPS)
   Tunable resistive pulse sensing (TRPS) was performed using the
   Nanocoulter G system (Resun Technology, China) equipped with a nanopore
   chip capable of detecting particles in the 50–250 nm size range. This
   technique was employed to measure both the concentration and size
   distribution of exosomes.
nanoLCMS/MS analysis
   nanoLCMS/MS analysis was performed for the identification of Munc13-4
   acetylation. 24 h following the co-expression of GFP-Munc13-4 and
   Flag-CBP 1323–1700 in HEK293T cells, GFP-Munc13-4 was
   immunoprecipitated from cell lysates using anti-GFP magnetic beads
   (Biolinkedin, L-1016), following the manufacturer’s protocol. The beads
   were subsequently washed three times with PBS, after which diluted 1×
   SDS-PAGE Sample Loading Buffer (Yeasen, 20315ES05) was added, and the
   samples were heated at 100 °C for 10 min. Proteins separated by
   SDS-PAGE were subjected to trypsin digestion (Promega, V5280) in 100 mM
   NH4HCO3 overnight at 37 °C. The resulting peptides were extracted with
   extraction buffer (1:2, vol/vol, 5% formic acid/acetonitrile) and then
   vacuum-dried.
   A total of 200 ng of peptides were separated and analyzed using a
   nano-UPLC system (Evosep One) coupled to a timsTOF Pro2 mass
   spectrometer (Bruker) equipped with a nano-electrospray ionization
   source. Peptide separation was achieved on a reversed-phase column
   (PePSep C18, 1.9 µm, 150 µm × 15 cm, Bruker) with mobile phases
   consisting of H2O containing 0.1% formic acid (phase A) and
   acetonitrile (ACN) with 0.1% formic acid (phase B). A 44-min gradient
   was used for the separation. Data acquisition was performed in DDA
   PaSEF mode, with the mass spectrometer scanning within a range of 100
   to 1700 m/z for MS1. During PASEF MS/MS acquisition, the collision
   energy was linearly increased in correlation with ion mobility, ranging
   from 20 eV (1/K0 = 0.6 Vs/cm²) to 59 eV (1/K0 = 1.6 Vs/cm²).
   The raw MS data files provided by the vendor were processed using
   SpectroMine software (version 4.2.230428.52329) in conjunction with the
   integrated Pulsar search engine. The MS spectra were queried against
   the species-specific UniProt FASTA database for Homo sapiens
   (uniprot_Homo sapiens_9606_reviewed_2023_09.fasta), with
   carbamidomethylation of cysteine (C) set as a fixed modification, and
   oxidation (M) and acetylation at the protein N-terminus as variable
   modifications. Trypsin was employed as the protease, with a maximum
   allowance of two missed cleavages. A false discovery rate (FDR) of 0.01
   was applied at both the peptide-spectrum match (PSM) and peptide
   levels. Peptide identification was performed with an initial precursor
   mass tolerance of 20 ppm. All other parameters were left at their
   default settings.
Proximity ligation assay (PLA)
   SUM159 cells were plated onto the wells of 29-mm glass-bottom dishes
   (Cellvis, D29-10-0-N). After fixation with 4% (wt/vol) paraformaldehyde
   at room temperature for 15 min, the cells were permeabilized with 0.2%
   (vol/vol) Triton X-100 in PBS for 10 min at room temperature, followed
   by three washes with PBS. Subsequently, in situ proximity ligation
   assays were performed using the Duolink^® In Situ Red Starter Kit
   Mouse/Rabbit (Sigma-Aldrich, DUO92101), following the manufacturer’s
   protocol. This included blocking, primary antibody incubation,
   Duolink^® PLA probe incubation, ligation, amplification, final washing,
   and nuclear staining in sequence. The cell samples were imaged using an
   FV3000 Confocal Laser Scanning Microscope (Olympus) equipped with a 60×
   oil-immersion objective (NA 1.42). DAPI was excited using a 405 nm
   laser, and Duolink^® In Situ Detection Reagents Red were excited with a
   594 nm laser. To determine the interaction between Munc13-4 and PD-L1,
   Munc13-4 Antibody (C-2) (Santa Cruz, sc-271300, 1:200) and PD-L1/CD274
   (C-terminal) Polyclonal antibody (Proteintech, 28076-1-AP, 1:200) were
   used. To explore the interaction between HRS and PD-L1, HGS Polyclonal
   antibody (Proteintech, 10390-1-AP, 1:500) and PD-L1/CD274 Monoclonal
   antibody (Proteintech, 66248-1-Ig, 1:200) were employed. Data were
   processed by home-written MATLAB script
   ([309]https://github.com/shenwang3333/PLA_Counting).
Orthotopic mouse models of breast cancer and treatments
   Mice were anesthetized via intraperitoneal injection of pentobarbital
   sodium at a dose of 40 mg/kg. Orthotopic breast cancer models were
   established by injecting control or Munc13-4 knockout 4T1 cells
   (3 × 10^5 cells per mouse) into the right fourth mammary fat pad of
   BALB/c, BALB/c Nude, or NOD/SCID female mice. Tumor dimensions were
   measured every other day starting on either day 4 or day 6
   post-inoculation using a digital caliper, and tumor volume was
   calculated using the formula: (width² × length × 0.5). At the end of
   the observation period, mice were euthanized, and tumors were
   harvested, weighed, and photographed for further analysis. According to
   the IACUC guidelines, the maximal allowable tumor size in adult mice is
   20 mm in diameter in any direction. In this study, the tumor
   size/burden in all experimental animals did not exceed this limit.
   Exosomes (1 × 10^9 particles) secreted by control 4T1 cells were
   pre-incubated with either an IgG isotype control antibody (Bioxcell,
   BE0090, 1:100) or InVivoMAb anti-mouse PD-L1 antibody (Bioxcell,
   BE0101, 1:100). To remove unbound antibodies, the exosome-antibody
   complexes were subjected to ultracentrifugation at 100,000 × g for
   70 min at 4 °C. The purified exosomes were then intravenously injected
   into mice via the tail vein. Injections were administered every other
   day for a total of nine treatments.
   Peptide treatments were performed using P-pep (Mus musculus) or S-pep,
   administered via intraperitoneal injection at a dosage of 100 µg per
   mouse. Once the tumor volume reached approximately 100 mm³, peptides
   were administered every other day, with a total of six injections.
   For immune checkpoint blockade (ICB) studies, mice were treated with
   IgG isotype control antibody (Bioxcell, BE0090, 100 µg/mouse),
   InVivoMAb anti-mouse PD-L1 antibody (Bioxcell, BE0101, 100 µg/mouse),
   or InVivoMAb anti-mouse PD-1 antibody (Bioxcell, BE0146, 100 µg/mouse).
   Antibodies were administered via intraperitoneal injection every 3 days
   for a total of six treatments.
Immune profiling
   To analyze T cell infiltration and activation, tumors, spleens, and
   tumor-draining lymph nodes (TDLNs) were excised from orthotopic mouse
   models of breast cancer. Tumor tissues were minced into small pieces
   and incubated with RPMI 1640 medium containing 1 mg/ml collagenase D
   (Roche, COLLD-RO) and 0.2 mg/ml DNase I (BioFroxx, 112MG010) at 37 °C
   for 1 h, followed by mechanical dissociation using a mesh cell
   strainer. The cells were then centrifuged at 500 × g for 5 min, washed
   with PBS, and treated with red blood cell (RBC) lysis buffer (BD
   Biosciences, 555899) to remove RBCs. The resulting cell suspension was
   filtered twice through a 70 μm nylon mesh to obtain single-cell
   suspensions. Lymphocytes from the spleens and TDLNs were isolated by
   mechanically squashing the tissues through a 70 μm mesh and removing
   RBCs.
   For stimulation, the single-cell suspensions were incubated with RPMI
   1640 medium containing Leukocyte Activation Cocktail (BD Biosciences,
   550583, 1:1000) at 37 °C for 6 h. After staining with viability dye FVS
   575V (BD Biosciences, 565694, 1:1000) to exclude dead cells, the cells
   were stained with the following antibodies. For surface marker
   analysis, cells were stained with anti-CD45-APC-Cy7 (BD Biosciences,
   557659, 1:50), anti-CD3-BV421 (BD Biosciences, 562600, 1:50),
   anti-CD4-Alexa Fluor 700 (BD Biosciences, 557956, 1:50), and
   anti-CD8-Percp-Cy5.5 (BD Biosciences, 551162, 1:50). For intracellular
   cytokine staining, cells were fixed and permeabilized using Fix/Perm
   buffer (BD Biosciences, 562574) and Perm/Wash buffer (BD Biosciences,
   562574), then re-stained with anti-KI67-BV510 (BD Biosciences, 563462,
   1:50), anti-IFN-γ-BV650 (BD Biosciences, 563854, 1:50), or
   anti-Granzyme B-FITC (Invitrogen, 11-8898-82, 1:500). Flow cytometric
   analysis was performed using the CytoFLEX flow cytometer.
CD8^+ T cell suppression assay
   To block PD-L1 on the exosome surface, equal quantities of exosomes
   isolated from control and Munc13-4 knockout 4T1 cells were incubated
   with PD-L1 blocking antibodies (Bioxcell, BE0101, 1:100) or IgG isotype
   control antibody (Bioxcell, BE0090, 1:100) at room temperature for 5 h.
   After incubation, the exosomes were washed with 25 ml PBS and subjected
   to ultracentrifugation to remove unbound antibodies.
   Mouse CD8^+ T cells were purified from splenocytes using the Mouse CD8
   T Cell Isolation Kit (Vazyme, CS103-01) and stimulated for 24 h with
   anti-CD3 (Biolegend, 300301, 1 μg/ml) and anti-CD28 (Biolegend, 117003,
   1 μg/ml) antibodies. Post-stimulation, the CD8⁺ T cells were incubated
   with the pre-processed exosomes for 16 h in the continued presence of
   anti-CD3 and anti-CD28 antibodies.
   Following treatment, CD8^+ T cells were harvested, stained with
   anti-CD8-Percp-Cy5.5 (BD Biosciences, 551162, 1:50), and permeabilized
   using Fix/Perm buffer and Perm/Wash buffer (BD Biosciences, 562574).
   Fixed cells were subsequently stained with anti-Granzyme B-FITC
   (Invitrogen, 11-8898-82, 1:500) and analyzed by flow cytometry.
T cell-mediated tumor cell killing assay
   Spleens were aseptically harvested from BALB/c mice and placed in
   sterile petri dishes containing cold PBS. The spleens were gently
   disrupted by grinding against a 70 μm cell strainer using a sterile
   syringe plunger, and the resulting cell suspension was centrifuged at
   500 × g for 5 min at 4 °C. The pellet was resuspended in 5 ml of PBS,
   and the suspension was carefully layered onto 5 ml of Ficoll-Paque™
   PLUS (Cytiva, 17144002) in a 15 ml conical tube without mixing. The
   gradient was centrifuged at 1000 × g for 20 min at room temperature
   with the centrifuge brake turned off. The mononuclear cell layer at the
   interface was carefully collected using a pipette, transferred to a
   fresh tube, and washed twice with PBS by centrifugation at 500 × g for
   5 min at 4 °C to remove residual Ficoll. The purified lymphocytes were
   then resuspended and stimulated with anti-CD3 (Biolegend, 300301,
   1 μg/ml) and anti-CD28 (Biolegend, 117003, 1 μg/ml) antibodies for
   24 h. Subsequently, the stimulated lymphocytes were co-cultured with
   adherent control or Munc13-4 knockout 4T1 cells in 96-well plates at an
   effector-to-target (E: T) ratio of 1:1 for 24 h. The viability of
   control and Munc13-4 knockout 4T1 cells was assessed using the Cell
   Counting Kit-8 (Vazyme, A311-01) following the manufacturer’s
   instructions.
Quantitative reverse transcription (qRT)-PCR assay
   Total RNA was extracted from each sample using the AFTSpin Tissue/Cell
   Fast RNA Extraction Kit for Animal (Abclonal, RK30120). The isolated
   RNA was eluted in nuclease-free water and reverse-transcribed into
   complementary DNA (cDNA) using the ABScript II cDNA First-Strand
   Synthesis Kit (Abclonal, RK20400). The resulting cDNA samples were
   subjected to quantitative PCR on a QuantStudio™ 6 Pro Real-Time PCR
   system using SYBR Green (Abclonal, RK21203) for detection.
   The primers used for qPCR were as follows:
   Hs-Munc13-4 qPCR forward primer: CCCTTTGTCCAGCTGACCTT
   Hs-Munc13-4 qPCR reverse primer: AGCAGGCACCAGGAATTCAA
   Hs-Actin qPCR forward primer: GCCGCCAGCTCACCAT
   Hs-Actin qPCR reverse primer: AGGAATCCTTCTGACCCATGC
Fluorescence imaging
   To analyze the colocalization between PD-L1 and CD63/LAMP1, Munc13-4
   knockout and control SUM159 cells were seeded onto glass-bottom dishes
   and transfected with the pEGFP-N3-PD-L1 plasmid along with either
   NPY-td-Orange2-CD63 or NPY-td-Orange2-LAMP1 plasmids. Following 24–36 h
   of transfection, cells were fixed with 4% paraformaldehyde (PFA) for
   15 min. For immunofluorescence staining, cells were permeabilized with
   0.2% Triton X-100 for 10 min and subsequently blocked with 5% bovine
   serum albumin (BSA) for 1 h at room temperature. After blocking, cells
   were incubated with primary antibodies overnight at 4 °C and washed
   three times with PBS. Secondary antibody staining was performed for 2 h
   at room temperature, followed by three PBS washes. Imaging was
   conducted using a Nikon confocal microscope equipped with a 60×
   oil-immersion objective lens (NA 1.40). The primary antibodies used in
   immunofluorescence experiments are as follows: CD63 Antibody
   (MX-49.129.5) (Santa Cruz, sc-5275, 1:100), LAMP1/CD107a Rabbit mAb
   (Abclonal, A21194, 1:200) and CBP/KAT3A/CREBBP Antibody (C-1) (Santa
   Cruz, sc-7300, 1:200). The secondary antibodies used in
   immunofluorescence experiments are as follows: Goat anti-Mouse IgG
   (H+L) Cross-Adsorbed Secondary Antibody, Alexa Fluor™ 488 (Invitrogen,
   A-11001, 1:500), Goat anti-Rabbit IgG (H+L) Highly Cross-Adsorbed
   Secondary Antibody, Alexa Fluor™ Plus 647 (Invitrogen, A-32733, 1:500).
   Fluorescence intensity was quantified using NIS-Elements AR 4.40
   software, and colocalization was assessed by calculating the Pearson’s
   correlation coefficient, also employing NIS-Elements AR 4.40 software.
TIRF microscopy for monitoring MVB mobility
   The mobility of exosomes in live cells was assessed using total
   internal reflection fluorescence (TIRF) microscopy (Nikon). To evaluate
   the effects of mutations in Munc13-4 on MVB mobility, control and
   Munc13-4 knockout SUM159 cells were seeded onto glass-bottom dishes.
   These cells were subsequently co-transfected with NPY-td-Orange2-CD63
   and pEGFP-N3-Munc13-4 WT or Munc13-4 mutants. Similarly, to examine the
   effects of Rab27a mutations on MVB mobility, control or Rab27a knockout
   SUM159 cells were plated and co-transfected with NPY-td-Orange2-CD63
   and pEGFP-C1-Rab27a WT or Rab27a mutants. After 24–48 h
   post-transfection, live-cell imaging was performed using a Nikon Ti
   inverted TIRF microscopy system equipped with a 100× oil-immersion
   objective (NA 1.49) and an EMCCD camera (Andor DU897). Orange
   fluorescence was excited using a 532 nm laser with an exposure time of
   300 ms. The diffusion coefficient (D), representing MVB mobility, was
   quantified using the Python-based trackpy library
   ([310]https://soft-matter.github.io/trackpy/dev/tutorial/walkthrough.ht
   ml).
Protein expression and purification
   Protein expression of human Munc13-4 and its mutants was performed
   using the Bac-to-Bac™ baculovirus expression system (Invitrogen).
   Briefly, recombinant bacmid DNA extracted from DH10Bac was transfected
   into Spodoptera frugiperda clone 9 (Sf9) cells using X-tremeGENE™ 9 DNA
   Transfection Reagent (Roche) to produce P1 baculovirus. Sequential
   infection of Sf9 cells with P1 baculovirus generated P2 and P3
   baculoviruses. For protein expression, Sf9 cells were infected with P3
   baculovirus and cultured for 48 h. Harvested cells were resuspended in
   lysis buffer (20 mM Tris-HCl, pH 8.1, 150 mM NaCl) supplemented with
   protease inhibitors (2 μg/ml aprotinin, 1 μg/ml leupeptin, 1 μg/ml
   pepstatin, and 1 mM PMSF). Cells were lysed using an AH-1500 Nano
   Homogenizer (ATS Engineering Inc.) at 800 bar under 4 °C. The lysate
   was clarified by centrifugation at 16,000 rpm using a JA-25.50 rotor
   (Beckman Coulter) at 4 °C. Supernatants were incubated with
   nickel-nitrilotriacetic acid (Ni-NTA) agarose (Qiagen) for 1 h at 4 °C,
   followed by two washes with wash buffer (20 mM Tris-HCl, pH 8.1, 150 mM
   NaCl, 20 mM imidazole). Bound proteins were eluted using buffer
   containing 20 mM Tris-HCl, pH 8.1, 150 mM NaCl, and 300 mM imidazole.
   Eluted proteins were further purified by size-exclusion chromatography
   using a Superdex™ 200 10/300 GL column (Cytiva).
   Other proteins were expressed in Escherichia coli BL21 (DE3). Protein
   expression was induced with isopropyl β-D-1-thiogalactopyranoside
   (IPTG). Harvested cells were lysed as described above. For GST-tagged
   proteins, clarified lysates were incubated with glutathione Sepharose
   4B (GE Healthcare), washed, and eluted using buffer containing 20 mM
   glutathione (neoFroxx, 1392GR025), 20 mM Tris-HCl, pH 8.1, and 150 mM
   NaCl. His-tagged proteins were purified as described for Munc13-4.
   Purified proteins were further processed using ion exchange and
   size-exclusion chromatography. For transmembrane proteins, 1.5% sodium
   deoxycholate was included throughout the purification process. Proteins
   were used immediately after affinity purification.
   For cryo-EM complex preparation, Munc13-4 and Rab27a were mixed at a
   molar ratio of 1:2 in the presence of 1 mM GppNHp (Aladdin, G276465)
   and incubated at room temperature for 1 h. The mixture was subjected to
   size-exclusion chromatography to isolate the stable protein complex.
GST pull-down assay
   For all GST pull-down assays, 2 μM of the GST-tagged protein was
   incubated with 3 μM of the target protein at room temperature for 2 h.
   Subsequently, the protein mixture was combined with glutathione
   Sepharose 4B resin (GE Healthcare) and incubated at 4 °C for 1 h. The
   resin was then washed four times with wash buffer (20 mM Tris-HCl, pH
   8.1, 150 mM NaCl, 0.02% Triton X-100) to remove unbound proteins. Bound
   proteins were eluted using an elution buffer containing 20 mM Tris-HCl
   (pH 8.1), 150 mM NaCl, 0.02% Triton X-100, and 50 mM glutathione.
   Eluted proteins were analyzed by SDS-PAGE.
SNARE assembly assay
   For SNARE assembly assay, VAMP-7 SNARE motif (A131C) was labeled with
   the Förster resonance energy transfer (FRET) donor dye BODIPY FL
   (Molecular Probes, [311]B10250), and SN-23-6CS (S161C) was labeled with
   the FRET-acceptor dye 5-tetramethylrhodamine (5-TAMRA) (Molecular
   Probes, T6027). During the experiment, both the donor protein and
   acceptor protein were at a concentration of 0.5 µM. The concentration
   of Syx-4 ΔTM (residues 1–275) was 2 µM, and Munc13-4 was at 10 µM. The
   experiments were performed using a PTI QM-40 spectrophotometer, with an
   excitation wavelength of 485 nm and an emission wavelength of
   513/580 nm, at room temperature. SNARE complex formation signals were
   interpreted as the FRET proximity ratio (E[PR]) between the donor
   (BODIPY) and acceptor (5-TAMRA). The E[PR] was determined using
   Eq. [312]1:
   [MATH:
   EPR=I5−
   TAMRA
   mrow>I5−T
   AMRA+IBODIPY :MATH]
   1
   where I[5-TAMRA] and I[BODIPY] represent the fluorescence intensities
   of 5-TAMRA and BODIPY FL, respectively, measured under the 485/10
   excitation filter.
Liposome fusion assay
   All lipids were dissolved at an initial concentration of 10 mg/ml in
   chloroform, except for PI(4,5)P₂, which was dissolved in a chloroform:
   methanol: water mixture (20:9:1) at 1 mg/ml. For Syx-4-liposome
   preparation, 52% 1-palmitoyl-2-oleoyl-glycero-3-phosphocholine (POPC;
   Avanti Polar Lipids, 850457), 20%
   1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoethanolamine (POPE; Avanti
   Polar Lipids, 850757), 15% 1,2-dioleoyl-sn-glycero-3-phospho-L-serine
   (DOPS; Avanti Polar Lipids, 840035), 10% cholesterol (Avanti Polar
   Lipids, 700000), 1% PI(4,5)P₂ (Avanti Polar Lipids, 840046), and 2% DiD
   (Molecular Probes, D307) were mixed to a final lipid concentration of
   1 mM. For VAMP-7-liposomes, 38% POPC, 11% POPE, 7%
   1,2-dioleoyl-sn-glycero-3-phospho-(1’-myo-inositol) (PI; Avanti Polar
   Lipids, 850149), 30% cholesterol, 15% sphingomyelin (Avanti Polar
   Lipids, 860584), and 3% DiI (Molecular Probes, D282) were mixed to the
   same final concentration. The lipid mixtures were vacuum-dried and
   resuspended in 2% sodium deoxycholate. Full-length Syx-4 protein (5 µM)
   and VAMP-7 SNARE-TM protein (5 µM) were incorporated into corresponding
   liposomes, respectively, by incubation at room temperature for 20 min.
   Detergent removal was performed using PD-10 desalting columns (GE
   Healthcare). The resulting liposomes were combined in a 1:1 ratio with
   a 60% (w/v) OptiPrep™ stock solution (Serumwerk Bernburg AG), layered
   with 2 ml of 20% (w/v) OptiPrep™ solution, and 200 µL of 20 mM Tris-HCl
   (pH 8.1), 150 mM NaCl. Liposomes were centrifuged at 120,000 × g for
   5 h at 18 °C using an SW55Ti rotor (Beckman). The top fraction was
   collected and dialyzed in a Slide-A-Lyzer™ dialysis cassette (Thermo
   Fisher, 66383) for 12 h before use. For the fusion assay, 10 µM SN-23
   was pre-incubated with Syx-4-liposomes at 37 °C for 2 h, while a
   negative control group was prepared without SN-23. Equal volumes of
   Syx-4- and VAMP-7-liposomes were mixed to a total volume of 60 µL, and
   0.4 µM Munc13-4 protein was added to the experimental group. Fusion was
   monitored using a FluoDia T70 fluorescence plate reader (Photon
   Technology Incorporated) at 37 °C, with excitation at 530 nm and
   emission at 580 nm and 667 nm. Liposome fusion signals were quantified
   by calculating the FRET proximity ratio (E[PR]) between the donor (DiI)
   and acceptor (DiD). The E[PR] was determined using Eq. [313]2:
   [MATH:
   EPR=IDi
   DI
   DiD+IDiI :MATH]
   2
   where I[DiD] and I[DiI] represent the fluorescence intensities of DiD
   and DiI, respectively, measured under the 530/10 excitation filter.
Co-flotation experiment of PD-L1-containing liposomes with Munc13-4
   For the preparation of PD-L1-containing liposomes, 80%
   1-palmitoyl-2-oleoyl-glycero-3-phosphocholine (POPC; Avanti Polar
   Lipids, 850457) and 20% 1,2-dioleoyl-sn-glycero-3-phospho-L-serine
   (DOPS; Avanti Polar Lipids, 840035) were combined to achieve a total
   lipid concentration of 1 mM. Lipid mixtures were vacuum-dried and
   resuspended in 2% sodium deoxycholate. PD-L1 protein (5 µM) was
   incorporated into the liposomes, followed by detergent removal using
   PD-10 desalting columns (GE Healthcare). The resulting liposome
   suspension (200 µL) was mixed in a 1:1 ratio with an 80% (w/v)
   Histodenz™ stock solution (Thermo Fisher, D2158). The mixture was
   layered sequentially with 350 µL of 30% (w/v) Histodenz™ and 20 µL of
   20 mM Tris-HCl (pH 8.1), 150 mM NaCl. The prepared gradient was
   centrifuged at 240,000 × g for 1.5 h at 18 °C using an SW55Ti rotor
   (Beckman). Fractions (20 µL) were sequentially collected from the top
   three layers of the gradient, with an additional 20 µL sample retrieved
   from the bottom layer. These fractions were analyzed by Western blot to
   assess the co-flotation of PD-L1-containing liposomes with Munc13-4.
Cryo-EM sample preparation
   The Munc13-4–Rab27a complex, bound to GppNHp, was prepared at a
   concentration of 0.5 mg/ml for cryo-EM analysis. Samples (3.5 µL) were
   applied to glow-discharged cryo-EM grids (Quantifoil, Cu, R1.2/R1.3,
   300 mesh) in an environment of 100% humidity at 4 °C. Grids were
   blotted for 2 s with a blotting force of 4 and subsequently vitrified
   by plunging into liquid ethane using a Vitrobot Mark IV (Thermo Fisher
   Scientific). Prepared grids were either screened immediately or stored
   in liquid nitrogen for future use.
Cryo-EM data acquisition and image processing
   The Munc13-4–Rab27a complex with GppNHp datasets were collected by
   300 kV Titan Krios electron microscope (Thermo Fisher Scientific)
   equipped with a Falcon4 direct electron detector coupled with a
   SelectrisX energy filter (10 eV slit width). The automated collection
   was performed using the EPU software in electron event representation
   (EER) mode, and all micrographs were recorded at a nominal
   magnification of 165,000× with a raw pixel size of 0.73 Å on the image
   plane. The micrographs were recorded in a −0.8 μm to −2.4 μm defocus
   range, with an electron dose rate of 11.47 e^– /Å^2 /s and a total dose
   of 50 e^– /Å^2. All the EER movies were pre-processed by CryoSPARC
   (version 4.5.3)^[314]68 to perform the motion correction and CTF
   estimation. 1,424,230 particles were selected by Blob Picking and
   subsequently subjected to three rounds of 2D classification, and
   317,449 particles from the selected classes were subjected to
   ab-initial 3D reconstruction. The initial volume was further refined by
   heterogeneous refinement and non-uniform refinement to yield a
   consensus map with 3.36 Å global resolution. To acquire a map with
   improved characteristics of the C2A and C2B domains, a total of 102,381
   particles with apparent features were selected from the 3D
   classification, yielding a reconstruction at 3.42 Å resolution.
   Particle subtraction was subsequently applied to these particles,
   focusing on the regions surrounding the C2A and C2B domains,
   respectively. Local refinement of the resulting datasets produced
   focused maps with resolutions of 4.39 Å and 7.2 Å. These two local
   volume maps were then combined during model building, producing a
   composite map with a resolution of 4.38 Å. All reported resolutions
   were estimated using the gold-standard Fourier shell correction 0.143
   criterion^[315]69. Data collection and refinement statistics are
   summarized in Supplementary Table [316]1.
Model building and refinement of the Munc13-4–Rab27a complex
   Initial models of Munc13-4 (AF-[317]Q70J99-F1) and Rab27a
   (AF-[318]P51159-F1) were generated using AlphaFold 2
   ([319]https://colab.research.google.com/github/deepmind/alphafold/blob/
   main/notebooks/AlphaFold.ipynb)^[320]70. The predicted structures were
   fitted into the cryo-EM density map through rigid-body docking using
   UCSF ChimeraX (v1.7.1)^[321]71. Further manual adjustments were
   performed using COOT (v0.9.6)^[322]72. Subsequent real-space refinement
   of the models was carried out through multiple iterative rounds using
   PHENIX (v1.20)^[323]73, followed by final model validation. Figures
   were prepared using PyMOL ([324]https://pymol.org/2/) and UCSF
   ChimeraX. Data validation statistics are summarized in Supplementary
   Table [325]1.
Statistics and reproducibility
   Statistical analyses were conducted using GraphPad Prism software
   (version 9.3). All key findings were reproduced in at least three
   independent experiments. Specific statistical tests employed for each
   experiment are detailed in the corresponding figure legends. For
   comparisons between two groups, either a two-tailed unpaired or paired
   t-test was applied, as appropriate. For multiple group comparisons,
   one-way analysis of variance (ANOVA) followed by Tukey’s multiple
   comparisons test, multiple t-tests, or two-way ANOVA with Sidak’s
   multiple comparisons test was utilized. A p-value of less than 0.05
   (P < 0.05) was considered indicative of statistical significance.
   Additional information on the study design, the number of replicates,
   and the statistical methods used is shown in the figure legends.
Reporting summary
   Further information on research design is available in the [326]Nature
   Portfolio Reporting Summary linked to this article.
Supplementary information
   [327]Supplementary Information^ (7.9MB, pdf)
   [328]Reporting Summary^ (5.2MB, pdf)
   [329]Transparent Peer Review file^ (1.4MB, pdf)
Source data
   [330]Source data^ (38.2MB, xlsx)
Acknowledgements