Abstract Dysregulation of the JAK2/STAT3 pathway disrupts immune balance and tissue homeostasis, leading to persistent inflammation and subsequent bone loss. Yet, commonly used JAK2 inhibitors with hydrophobicity primarily suppress inflammation but often fail to provide sustained pharmacological effects and address ongoing inflammation-induced bone destruction. Here, using computer-aided drug design, we developed hydrophilic calcium-doped carbon dots (Ca-CDs) as a multifunctional nanoinhibitor targeting JAK2. The Ca-CDs can block the excessive activation of the JAK2/STAT3 pathway by binding to JAK2, thereby reducing the secretion of pro-inflammatory cytokines and exerting anti-inflammatory effects. Furthermore, the Ca-CDs promote bone regeneration under inflammatory conditions and serve as crosslinking junctions that facilitate the formation of a Ca-CDs-based alginate hydrogel, thereby enabling prolonged drug retention and controlled release. Their application in the representative inflammatory microenvironment of periodontitis successfully validated the efficacy of the Ca-CDs-based therapeutic strategy. Overall, targeting JAK2 with Ca-CDs nanoinhibitor offers a promising treatment option for JAK2-associated inflammatory diseases. Graphical Abstract [50]graphic file with name 12951_2025_3744_Figa_HTML.jpg Supplementary Information The online version contains supplementary material available at 10.1186/s12951-025-03744-w. Keywords: Carbon dots, JAK2 nanoinhibitor, Sustained release, Anti-inflammation, Bone repair Introduction Janus protein tyrosine kinases (JAK1, JAK2, JAK3, and TYK2) are crucial mediators of the Janus kinase/signal transducer and activator of transcription (JAK/STAT) pathway, essential for transmitting extracellular cytokine-mediated signals across the cell membrane to the nucleus, ultimately regulating gene transcription [[51]1–[52]3]. This pathway is pivotal for maintaining hematopoiesis, immune balance, and tissue homeostasis. Dysregulation of JAK/STAT signaling is closely associated with inflammatory and immune diseases such as rheumatoid arthritis, periodontitis, and inflammatory bowel disease [[53]4–[54]6]. In particular, the overactivation of the cytokine-mediated JAK2/STAT3 pathway induces the expansion and persistence of inflammation, which in turn causes the continuous destruction of bone in these diseases [[55]7]. Given its critical role in human health, targeting this pathway presents a significant therapeutic opportunity for managing a broad spectrum of conditions [[56]8]. Currently, commonly used JAK2 inhibitors are small molecules characterized by NH/O spacers, amide bonds, and N-heterocyclic structures [[57]9–[58]11]. These inhibitors are typically administered orally due to their poor water solubility, which presents several limitations, including side effects and low therapeutic concentrations at target sites [[59]12–[60]14]. This issue is particularly problematic for inflammatory diseases requiring localized treatment, such as periodontitis, where hydrophobic inhibitors fail to remain at the lesion site, thus diminishing their effectiveness [[61]15]. More importantly, these inhibitors lack multifunctionality and fail to address the ongoing bone destruction driven by inflammation. Therefore, to overcome these challenges, there is a significant need to design an inhibitor that not only more effectively regulates the JAK2/STAT3 pathway to suppress the secretion of pro-inflammatory cytokines but also prevents bone loss and promotes bone regeneration associated with inflammation [[62]16]. Carbon dots (CDs) have garnered extensive attention due to their unique advantages [[63]17–[64]19], such as the ability to engineer various functional groups by selecting specific chemical structures as precursors [[65]20–[66]22], potentially allowing direct involvement in cellular signaling regulation. In this study, leveraging the structures of JAK2 protein and known inhibitors, pharmacophore models were constructed using the HipHop algorithm, and potential fragments that fit well within the JAK2 binding site were identified through fragment-based drug discovery. Consequently, calcium-doped CDs (Ca-CDs), a multifunctional nanoinhibitor targeting JAK2, were synthesized through space-confined vacuum heating of acetylsalicylic acid, metformin, and calcium chloride. The synthesized Ca-CDs have been demonstrated to regulate the JAK2/STAT3 pathway effectively, thereby inhibiting the secretion of pro-inflammatory cytokines. The capability of Ca-CDs to control the JAK2/STAT3 pathway mainly stems from their efficient binding to JAK2 protein, which was elucidated through molecular dynamics simulation and surface plasmon resonance (SPR) analysis, exhibiting an affinity dissociation constant (K[D]) of 64.2 µM. Additionally, the calcium incorporated within the Ca-CDs facilitates the repair of bone defects caused by inflammation. Significantly, the inherent calcium in the Ca-CDs can trigger the formation of a Ca-CDs-based alginate (Ca-CDs@ALG) hydrogel, where Ca-CDs serve as crosslinking junctions, thereby enabling long-term retention at local lesion sites and sustained drug release. To validate the efficacy of Ca-CDs, periodontitis, a condition requiring local administration and prevalent in moist oral environments, was selected as a model. In vivo experiments demonstrated that our nanoinhibitor could effectively reduce the inflammatory response and promote the regeneration of bone defects caused by inflammation, showcasing its substantial potential for clinical applications in JAK2-associated inflammatory diseases (Fig. [67]1). Fig. 1. [68]Fig. 1 [69]Open in a new tab Schematic illustration of the design strategy for Ca-CDs targeting JAK2 and their anti-inflammatory and osteogenic mechanisms for the sustained treatment of inflammatory diseases. Ca-CDs, computer-aided designed as a multifunctional nanoinhibitor targeting the JAK2/STAT3 pathway, reduce pro-inflammatory cytokines secretion and promote osteogenic differentiation. The Ca-CDs also serve as crosslinking junctions to form the Ca-CDs@ALG hydrogel, enhancing therapeutic efficacy in vivo Results Computer-aided drug design, synthesis and characterization of Ca-CDs A chemically diverse set of 53 known JAK2 inhibitors was selected as the training set to generate a ligand-based common feature pharmacophore model using the HipHop algorithm. The selected compounds displayed key features such as hydrogen bond acceptors (HBA), hydrogen bond donors (HBD), ring aromaticity (RA), and hydrophobic groups (HYD) associated with their aromatic structures (Fig. S1). Additionally, all training set compounds shared an essential nitrogen-containing heterocyclic functionality. After calculations, ten pharmacophore models were generated, and the top two were shown based on their rank values (Fig. [70]2A and Video S1). Analysis of the resulting pharmacophore models revealed that, besides the presence of HBA, HBD, RA, and HYD features, HBA and RA were adjacent in pharmacophore model 1, while HBD and RA were adjacent in pharmacophore model 2, suggesting the presence of aromatic rings containing heteroatoms. Furthermore, fragment-based drug design was performed using the structures of the 53 inhibitors and the JAK2 protein from the UniProt database (ID: [71]Q62120) (Fig. [72]2B and Video S2). The results indicated that suitable fragments for interacting with JAK2 primarily contained nitrogen and/or oxygen aromatic rings (Fig. [73]2C). Therefore, the aforementioned pharmacophore model and fragment-based drug design can serve as a preliminary screening strategy, guiding the selection of precursors capable of forming such heterocyclic structures for CDs synthesis, thereby greatly reducing the workload of experimental validation. Fig. 2. [74]Fig. 2 [75]Open in a new tab Computer-aided drug design and characterizations of Ca-CDs. (A) Top 2 ranked pharmacophore models, highlighting key structural features: HBA (green), HBD (pink), RA (orange), and HYD aromatic (blue). (B) A general view of the three-dimensional structure of JAK2, shown as a rainbow-colored cartoon (left panels) and a gray-shaded surface model (right panels). A close-up highlights its binding site, represented by a purple sphere. (C) Five fragments suitable for interacting with the JAK2 binding site (upper panels) and their structures formulas (lower panels). JAK2 is depicted as a rainbow-colored cartoon and gray-shaded surface. The fragments are shown in ball-and-stick mode, with carbon, hydrogen, oxygen, and nitrogen atoms colored gray, white, red, and blue, respectively. (D) Schematic diagram of Ca-CDs synthesis (ASA: acetylsalicylic acid). Photographs of (E) Ca-CDs sphere and (F) Ca-CDs solution taken under sunlight (upper panels) and UV light (lower panels), respectively. (G) Excitation-emission map of Ca-CDs. (H) HRTEM image of Ca-CDs, with insets showing the corresponding particle size distribution histogram and an enlarged view of the Ca-CDs lattice. (I) FTIR spectra of acetylsalicylic acid (purple), metformin (blue), and Ca-CDs (red). The enlarged image shows the peak fitting of the FTIR spectra of Ca-CDs in the 1600–1800 cm^−1 range. High-resolution XPS spectra of Ca-CDs: (J) C 1 s and (K) N 1s. (L) EDS pattern of Ca-CDs. Inset: the flame test image of Ca-CDs solution Based on the above analysis results, multiple compounds containing nitrogen and/or oxygen aromatic rings from our drug library were selected as candidate precursors for CD synthesis and pharmacological evaluation (Fig. [76]S2). Consequently, acetylsalicylic acid and metformin were selected as organic precursors for preparing CDs. Using our previously developed space-confined vacuum heating method [[77]23], a solution containing acetylsalicylic acid, metformin, and calcium chloride was heated and evacuated (Fig. [78]2D), resulting in the formation of a foamy, porous spherical structure (Fig. [79]S3). Continued heating of the reaction system yielded a porous yellow spherical product that emitted bright fluorescence under ultraviolet (UV) excitation (Fig. [80]2E). After dialysis and purification, a light yellow Ca-CDs solution was obtained (Fig. [81]2F). The ultraviolet-visible (UV-vis) absorption spectrum and photoluminescence (PL) spectrum of this solution differed from those of the raw materials (Fig. [82]S4 and S5), confirming the successful synthesis of Ca-CDs. Additionally, the excitation-emission map presented a single emission center (Fig. [83]2G), demonstrating size uniformity, further confirmed by high-resolution transmission electron microscopy (HRTEM) (Fig. [84]2H). The HRTEM image revealed that the Ca-CDs are uniformly sized, with an average diameter of 3.5 nm. The well-defined lattice spacing of 0.24 nm corresponds to the (110) crystallographic planes of graphitic carbon, confirming the formation of a graphene-like, conjugated aromatic structure. To further clarify the structure of the Ca-CDs, Fourier transform infrared (FTIR) spectroscopy and X-ray photoelectron spectroscopy (XPS) characterization were performed. As depicted in Fig. [85]2I, the FTIR spectrum of acetylsalicylic acid reveals the presence of carboxyl group, evidenced by a C = O stretching vibration at 1690 cm^−1. The peaks at 1458 and.1605 cm^−1 indicate the presence of the carbon framework in benzene ring, which provides the structural basis for the formation of conjugated structures in Ca-CDs. The FTIR spectrum of metformin displays characteristics of amine group, confirmed through an N-H stretching vibration between 3040 and 3450 cm^−1, and an N-H bending vibration at 1570 cm^−1. Within the FTIR spectrum of Ca-CDs, a broad band peak ranging from 2800 to 3500 cm^−1 is attributed to O-H/N-H stretching vibrations. The peaks at 1654 cm^−1 and 1558 cm^−1 correspond to C = O stretching vibrations and N-H bending vibrations, respectively, which are associated with the amide I and amide II bands. These findings indicate the presence of amide bonds in Ca-CDs, formed through dehydration condensation between carboxyl and amine groups in the raw materials. Additionally, unsaturated carbon-nitrogen bond stretching vibrations are observed at 1631 and 2183 cm^−1, with C-N stretching vibration also detected at 1366 cm^−1, confirming the presence of arylamine [[86]24]. Furthermore, peaks at 1611 and 1466 cm^−1 are associated with the carbon framework (i.e., C = C stretching vibrations) in conjugated structures. The presence of numerous functional groups within the Ca-CDs is further substantiated by XPS characterizations. The C 1 s spectrum of acetylsalicylic acid showed four peaks at 284.7, 285.1, 286.3, and 288.8 eV, indicating the presence of sp^2 C, sp^3 C, C-O, and COO, respectively (Fig. S6A). The C 1 s spectrum of metformin displays three peaks at 284.9, 285.9, and 288.3 eV, indicative of sp^3 C, C-N, and guanidino groups (Fig. S6B). The N 1 s spectrum of metformin reveals three peaks at 398.4, 399.4, and 400.2 eV, corresponding to N-H, C-N, and C = N, respectively (Fig. S7). In comparison, as shown in Fig. [87]2J, K, the C 1 s spectrum of the Ca-CDs shows six peaks at 284.6, 285.1, 285.8, 286.4, 287.6, and 288.9 eV, corresponding to sp^2 C, sp^3 C, C-N, C-O, CON, and COO/guanidino groups, respectively (Fig. [88]2J). The N 1 s spectrum of Ca-CDs displays four peaks at 398.4, 399.5, 400.3, and 402.3 eV, representing N-H, pyrrolic N, pyridinic N, and graphitic N/CON, respectively (Fig. [89]2K). These findings underscore the amide formation and the presence of a highly conjugated sp^2 domain and heteroatom-containing rings in Ca-CDs, setting the stage for specific interactions with target protein. Energy-dispersive spectroscopy (EDS) characterization confirmed the presence of C, N, O, and Ca elements in the Ca-CDs, with the atomic ratio of Ca being 9.1% (Fig. [90]2L). To further verify the presence of calcium visually, a flame test was conducted, resulting in a brick-red flame, which is characteristic of calcium (Fig. [91]2L and S8; Video S3). The presence of calcium provides favorable conditions for the regeneration of bone damaged by inflammation. Biocompatibility and anti-inflammatory effects of Ca-CDs Before applying to in vivo and in vitro studies, the biocompatibility of Ca-CDs should be assessed. A cell counting kit-8 (CCK-8) was utilized to evaluate the cytotoxicity of RAW264.7 cells treated with Ca-CDs for 24 h. The results showed that Ca-CDs had no significant effect on the viability of RAW264.7 cells within the concentration range of 0–300 µg/mL. However, the viability decreased to 84.7% at a concentration of 400 µg/mL (Fig. [92]3A). Flow cytometry and live/dead cell staining further demonstrated that there were no significant changes in apoptosis and morphology of RAW264.7 cells at working concentrations ranging from 0 to 300 µg/mL (Fig. [93]3B and S9). These findings suggest that Ca-CDs have good biocompatibility. To directly observe the distribution of Ca-CDs within cells, the unique PL properties of Ca-CDs were employed for cell imaging using confocal laser scanning microscopy (CLSM). As shown in Fig. [94]3C, the blue fluorescence of Ca-CDs was primarily observed in the cytoplasm, indicating that the Ca-CDs can be uptaken and internalized by the RAW264.7 cells. This establishes a proof of concept for the pharmacological functionality and further research both in vivo and in vitro of Ca-CDs. Fig. 3. [95]Fig. 3 [96]Open in a new tab Ca-CDs exert anti-inflammatory effects via regulating JAK2/STAT3 pathway. (A) CCK-8 assay of RAW264.7 cells, which are treated with different concentrations of Ca-CDs (0, 50, 100, 200, 300, and 400 µg/mL). (B) Cell apoptosis assay of RAW264.7 cells, which are treated with different concentrations of Ca-CDs (0, 100, 200, and 300 µg/mL), respectively. (C) Cell imaging of Ca-CDs-treated RAW264.7 cells. (D) Detection of mRNA levels of Nos2, Il1b, and Tnfa by RT-qPCR. (E) Western blot analysis of pro-inflammatory cytokines: iNOS, IL-1β, and TNF-α. (F) Corresponding quantitative data from Western blot. (G) Volcano diagram (q-value < 0.05 & |log[2]FC|>1) and (H) heat map of differentially expressed genes. (I) Bubble diagram of KEGG pathway enrichment analysis. (J) GSEA enrichment plot of JAK/STAT pathway. (K) Western blot results of p-STAT3, STAT3, IL-1β, and TNF-α treated without or with LPS, Ca-CDs (300 µg/mL), and Colivelin (500 nM), respectively. (L) Western blot results of p-JAK2, JAK2, p-STAT3, STAT3, IL-1β, and TNF-α treated without or with LPS, Ca-CDs (300 µg/mL), and Coumermycin A1 (CA1, 10 µM), respectively. (M) Schematic diagram of Ca-CDs inhibiting pro-inflammatory cytokines secretion via regulating JAK2/STAT3 pathway. Data are presented as mean ± SD from three independent experiments. * indicates P < 0.05, ** indicates P < 0.01, *** indicates P < 0.001 Considering that Ca-CDs exhibit structural similarities in their functional groups with JAK2 inhibitors, it is anticipated that Ca-CDs may possess anti-inflammatory effects. To investigate the effects of Ca-CDs on the secretion of pro-inflammatory cytokines in RAW264.7 cells, lipopolysaccharide (LPS) was used to establish an inflammatory model in vitro. The cells were then co-incubated with Ca-CDs and another group received acetylsalicylic acid and metformin (ASA + MET) for 24 h, respectively. Reverse-transcriptase quantitative polymerase chain reaction (RT-qPCR) results indicated that compared to the LPS group, Ca-CDs significantly reduced the mRNA expression of Nos2, Il1b, and Tnfa, which are representative pro-inflammatory cytokines, as the concentration of Ca-CDs increased from 0 to 300 µg/mL (Fig. [97]3D). Western blot analysis further confirmed a significant reduction in the expression of iNOS, IL-1β, and TNF-α when treated with Ca-CDs (Fig. [98]3E, F). Moreover, these reductions were more pronounced than those observed in the ASA + MET group. This suggests that Ca-CDs possess potent pharmacological activity. Given that Ca-CDs are explored as a JAK2 nanoinhibitor, their anti-inflammatory effects were compared with those of Fedratinib, a clinically approved JAK2 inhibitor. At biosafe concentrations, RT-qPCR analysis revealed that our Ca-CDs exhibited anti-inflammatory activity comparable to that of the commercially available Fedratinib (Fig. S10), demonstrating the potential of Ca-CDs as a JAK2 nanoinhibitor. Ca-CDs inhibit the JAK2/STAT3 pathway to exert anti-inflammatory effects Having established the anti-inflammatory effects of Ca-CDs, RNA transcriptome analysis was conducted to explore the underlying mechanisms of this biological phenomenon. Compared to the LPS group, the addition of Ca-CDs to LPS-induced RAW264.7 cells resulted in the identification of 97 differentially expressed genes, including 12 upregulated and 85 downregulated genes (Fig. [99]3G). As depicted in Fig. [100]3H, the heat map shows significant downregulation of cytokines and chemokines receptors related to inflammation, such as Il6, Il23a, Lif, Il1a, Il1b, Cxcr5, and Ccr7. Further analysis through Gene Ontology (GO) enrichment and Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway enrichment of these differentially expressed genes aimed to identify the main biological functions and pathways involved. It was found that the Ca-CDs primarily participate in immune and inflammatory responses and are predominantly enriched in cytokine-related pathways, particularly the JAK/STAT pathway (Fig. [101]3I and S11). Gene Set Enrichment Analysis (GSEA) revealed substantial activation of the JAK/STAT pathway following LPS stimulation, which was effectively inhibited by Ca-CDs (Fig. [102]3J). Given the heterocyclic structures in Ca-CDs that are similar to those of JAK2 inhibitors, it is hypothesized that Ca-CDs may exert their anti-inflammatory effects by inhibiting the JAK2/STAT3 pathway. To validate this hypothesis, Western blot experiments were conducted. We first examined whether the expression and activation of the STAT3 in the JAK2/STAT3 pathway could be regulated by Ca-CDs. The results indicated that, compared to the control group, the phosphorylation level of STAT3 was significantly increased in the LPS group, confirming STAT3 activation during the inflammatory process (Fig. [103]3K and S12). However, after treating LPS-induced cells with Ca-CDs, both the phosphorylation level of STAT3 and pro-inflammatory cytokines expression were significantly reduced. To further explore whether the anti-inflammatory effects of Ca-CDs could be reversed, Colivelin, an STAT3 activator, was used [[104]25]. As shown in Fig. [105]3K and S12, when Colivelin was added to the LPS group, there were no significant variations in the phosphorylation level of STAT3, nor were there changes in the expression of IL-1β and TNF-α. However, when Colivelin and Ca-CDs were simultaneously administered to LPS-induced cells, there was a noticeable increase in the phosphorylation level of STAT3 compared to the LPS-induced cells treated solely with Ca-CDs, along with elevated expression of IL-1β and TNF-α. These results imply that the downregulation of pro-inflammatory cytokines by Ca-CDs is achieved through the inhibition of phosphorylation of STAT3. To investigate whether the inhibition of STAT3 phosphorylation is due to the upstream blockade of JAK2 by Ca-CDs, Coumermycin A1, a JAK2 activator, was used [[106]26]. As shown in Fig. [107]3L and S13, compared to the control group, there was a significant increase in the phosphorylation levels of JAK2 and STAT3 in the LPS group, indicating activation of the JAK2/STAT3 pathway. However, the phosphorylation levels of JAK2 and STAT3, as well as the expression of IL-1β and TNF-α, were significantly suppressed after treating LPS-induced cells with Ca-CDs. Additionally, when Coumermycin A1 and Ca-CDs were simultaneously administered to LPS-induced cells, the phosphorylation levels of JAK2 and STAT3 were essentially restored, and the expression of IL-1β and TNF-α increased accordingly. JAK2, an intracellular protein, comprises seven conserved domains without transmembrane regions. The carboxy-terminal domain possesses tyrosine kinase activity, which activates downstream signaling [[108]27]. As shown previously in Fig. [109]3C, Ca-CDs are predominantly distributed in the cytoplasm, providing a precondition for the interaction of Ca-CDs with the cytoplasmic domains of JAK2. Moreover, amino acid residues surrounding the JAK2 binding site typically interact with nitrogen-containing five- or six-membered heterocyclic compounds, thereby blocking the pathway [[110]28]. Based on previous HRTEM, FTIR, and XPS analyses, which confirmed that Ca-CDs possess a graphene-like carbonized core and a rich nitrogen-containing heterocyclic structure, it is reasonable to infer that Ca-CDs can interact with JAK2 to form a stable complex, thus reducing the phosphorylation level of JAK2, inhibiting the JAK2/STAT3 pathway and finally exerting anti-inflammatory effects (Fig. [111]3M). Ca-CDs inhibit the activation of JAK2/STAT3 pathway by binding to JAK2 Molecular dynamics simulations were used to model the binding process between Ca-CDs and JAK2 and to assess the structural stability. Before the simulation, an initial model was constructed using the JAK2 protein structure from the UniProt database, and a graphene-like fragment was attached to the ligand molecule as a simplified model of Ca-CDs (Fig. [112]4A). Since the JAK2 protein structure was predicted using AlphaFold, a Ramachandran plot was applied to verify the predicted torsion angles. As shown in Fig. S14, 98.5% of the residues in the protein were located in the allowed regions, demonstrating a high-quality model. A 120 ns molecular dynamics simulation was then performed. An animated movie of the JAK2/Ca-CDs complex, based on 600 frames extracted from the trajectory, was provided in Video S4. The movie and snapshot images of the simulation trajectory at 0, 60, and 120 ns showed that the Ca-CDs were stably bound to the JAK2 throughout the simulation process (Fig. [113]4A-C). Fig. 4. [114]Fig. 4 [115]Open in a new tab Ca-CDs inhibit the activation of JAK2/STAT3 pathway by binding to JAK2. (A-C) Structures of Ca-CDs binding to JAK2, extracted from the molecular dynamics simulation trajectory at (A) 0 ns, (B) 60 ns, and (C) 120 ns, respectively. JAK2 is depicted as a rainbow-colored cartoon and gray-shaded surface. Ca-CDs are shown in ball-and-stick mode, with carbon, hydrogen, oxygen, and nitrogen atoms colored gray, white, red, and blue, respectively. Enlarged views highlight the binding conformations of Ca-CDs within the JAK2 binding pocket at different simulation times. (D) RMSD of the Cα atoms of JAK2 in JAK2/Ca-CDs complex as a function of time. (E) Numbers of residues in diverse secondary structures of JAK2. (F) RMSF values of amino acid residues 750–1130 in JAK2/Ca-CDs complex and JAK2, respectively. (G) A schematic representation of SPR analysis illustrating the interaction between Ca-CDs and JAK2. JAK2 protein is covalently immobilized on a CM5 sensor chip as the stationary phase, with Ca-CDs introduced as the mobile phase. The binding of Ca-CDs to JAK2 causes a shift in the reflected angle. (H) The sensorgram of Ca-CDs binding to JAK2-immobilized sensor chip. (I) The equilibrium responses versus Ca-CDs concentrations A least-squares fit of the entire trajectory was conducted to calculate the root mean square deviation (RMSD) of the backbone atoms from their initial structure. The RMSD values stabilized at approximately 6 Å after 90 ns of simulation, indicating that the simulation reached stability beyond this point (Fig. [116]4D). Additionally, the number of residues in various secondary structures of the JAK2, including α-helix, β-sheets, turns, coil, and bend, remained nearly constant throughout the simulation (Fig. [117]4E). Further analysis involved measuring the residue-wise root mean square fluctuation (RMSF), which reflects deviations in the positions of JAK2 amino acid residues. Before evaluating the RMSF of the complex, a 120 ns molecular dynamics simulation of JAK2 without Ca-CDs was performed (Fig. S15A-C), with the RMSF values of JAK2 serving as a reference for comparison. The RMSD results indicated that JAK2 reached a stable conformation after 100 ns of simulation (Fig. S15D). Then, to calculate the RMSF, the simulation trajectories of both the JAK2/Ca-CDs complex and JAK2 alone were obtained from the last stable 20 ns of the molecular dynamics simulation. As shown in Fig. [118]4F, the RMSF values near the binding pocket of the JAK2/Ca-CDs complex were significantly lower than those of JAK2 alone, indicating that these amino acid residues were restrained, suggesting several interactions between Ca-CDs and JAK2. To further investigate the interaction between Ca-CDs and JAK2, we employed surface plasmon resonance (SPR) technology, recognized as the gold standard for real-time detection of molecular interactions [[119]29]. Purified recombinant JAK2 protein was covalently immobilized on a CM5 sensor chip to serve as the stationary phase for protein-ligand binding in the SPR analysis. Ca-CDs were introduced over the chip as the mobile phase. The binding of Ca-CDs to JAK2 led to real-time changes in the refractive index at the sensor surface, causing shifts in the resonance angle. By monitoring these changes, we obtained kinetic parameters and the equilibrium dissociation constant (K[D]) of the interaction (Fig. [120]4G). The results showed that the K[D] value between Ca-CDs and JAK2 was approximately 64.2 µM, exhibiting rapid association and dissociation kinetics (Fig. [121]4H, I). These findings confirm a significant interaction between Ca-CDs and JAK2, which contribute to the regulation of inflammatory pathways and thereby inhibit the expression of pro-inflammatory cytokines. Promotion of Ca-CDs on the osteogenic differentiation Dysregulation of the JAK2/STAT3 pathway disrupts immune homeostasis, leading to the expansion and persistence of inflammation, which subsequently contributes to continuous bone destruction. Consequently, the ultimate objective of treatment is to reshape the inflammatory microenvironment and reconstruct the bone tissue [[122]30]. To this end, our Ca-CDs show promise for promoting bone regeneration under inflammatory conditions by leveraging the incorporated calcium element (Fig. [123]5A). Initially, rat bone marrow stromal cells (rBMSCs) were isolated and identified (Fig. S16), followed by biosafety testing. The results demonstrated that Ca-CDs exhibited good biocompatibility within the concentration range of 0–300 µg/mL (Fig. [124]5B and S17-19) and were successfully taken up into the cytoplasm (Fig. [125]5C), providing a prerequisite condition for the biological functions of Ca-CDs in rBMSCs. Fig. 5. [126]Fig. 5 [127]Open in a new tab Ca-CDs promote the osteogenic differentiation under inflammatory conditions. (A) Schematic diagram of Ca-CDs protection for rBMSCs under LPS stimulation. (B) CCK-8 assay of rBMSCs, which are treated with different concentrations of Ca-CDs (0, 50, 100, 200, 300, and 400 µg/mL). (C) Cell imaging of Ca-CDs-treated rBMSCs. ALP staining on (D) day 7 and (E) day 14. (F) ARS staining on day 21. (G) Validation of osteogenic genes expression by RT-qPCR on day 7. (H) Western blot analysis of osteogenic-related proteins: RUNX2, SP7, and OCN. (I) Corresponding quantitative data from Western blot. (J) IF staining of OCN using CLSM. Data are presented as mean ± SD from three independent experiments. * indicates P < 0.05, ** indicates P < 0.01, *** indicates P < 0.001 Based on the biosafety results, the dosing concentration range for rBMSCs was set to match that of RAW264.7 cells. Initially, the effect of Ca-CDs on rBMSCs under normal conditions was examined, confirming their ability to promote osteogenic differentiation (Fig. S20-23), whereas Fedratinib, a JAK2 inhibitor, did not upregulate osteogenic gene expression (Fig. S22). These results indicate that, in contrast to a conventional JAK2 inhibitor, Ca-CDs possess additional osteogenic activity, highlighting their multifunctionality as a nanoinhibitor. Subsequently, LPS was used to stimulate rBMSCs to mimic an inflammatory microenvironment, followed by alkaline phosphatase (ALP) and alizarin red S (ARS) staining. The staining results showed that under LPS stimulation, almost no positive cells were observed, whereas the addition of Ca-CDs significantly increased both the number of positive cells and the staining intensity in a dose-dependent manner (Fig. [128]5D-F). RT-qPCR and Western blot results further demonstrated that Ca-CDs promoted the expression of osteogenic markers at both the genetic and protein levels (Fig. [129]5G-I). Immunofluorescence (IF) staining was further used to assess the expression of osteocalcin (OCN), a specific biochemical marker of bone formation. The results revealed that the intensity of red fluorescence from OCN in the Ca-CDs group was significantly higher than in the LPS group, indicating increased expression (Fig. [130]5J). These comprehensive experimental findings confirm that Ca-CDs effectively promote osteogenic differentiation in rBMSCs under both inflammatory and normal conditions. To further investigate whether the osteogenic effect of Ca-CDs is associated with the incorporated calcium, undoped CDs were synthesized from ASA and MET. RT-qPCR and ALP staining results showed that undoped CDs exhibited negligible osteogenic activity (Fig. S22 and S24), directly confirming that the osteogenic function of Ca-CDs depends on calcium doping. To clarify the contribution of calcium in Ca-CDs, the osteogenic effect of Ca-CDs were compared with those of calcium chloride (CaCl[2]) at an equivalent calcium concentration. The results demonstrated that Ca-CDs induced stronger osteogenic activity than CaCl[2] (Fig. S22 and S24), suggesting that the osteogenic function of Ca-CDs is not solely attributable to calcium release. The difference between Ca-CDs and CaCl[2] is most likely due to the nanoscale size of Ca-CDs, which facilitates enhanced cellular uptake and consequently improves Ca^2+ bioavailability. To verify this hypothesis, Rhod-2, a calcium-sensitive fluorescent probe, was used to measure intracellular calcium levels in rBMSCs (Fig. S25). The results showed that Ca-CDs significantly increased intracellular Ca^2+ levels compared with CaCl[2], thereby enhancing osteogenic effects. These findings confirm that the osteogenic effect of Ca-CDs is achieved through both calcium doping and enhanced cellular calcium uptake. Construction of Ca-CDs-based self-triggered hydrogel delivery system It is well known that reshaping the inflammatory microenvironment and reconstructing bone tissue is a slow and complex process requiring long-term regulation [[131]31]. However, maintaining prolonged drug residence at the target site remains challenging during treatment, limiting sustained pharmacological effects. Hydrogel-based delivery systems offer a promising solution for addressing these challenges in vivo [[132]32–[133]34]. Notably, the inherent calcium in the Ca-CDs can serve as crosslinking junctions by coordinating with carboxyl group, thereby triggering the formation of a Ca-CDs@ALG hydrogel (Fig. [134]6A). As demonstrated, the blue fluorescence of the formed hydrogel confirms the uniform distribution of Ca-CDs within the matrix (Fig. S26). The Ca-CDs@ALG hydrogel exhibits both injectable and moldable properties (Video S5). As shown in Fig. [135]6B, C and S27, the hydrogel can be injected to form the letters “JLU” and molded into a cylindrical shape. Furthermore, rheological testing was conducted to assess the mechanical properties of the hydrogel. As the angular frequency increased, the storage modulus (G’) remained consistently higher than the loss modulus (G’’) (Fig. [136]6D), indicating that the hydrogel can withstand rapid deformation and maintain a stable network structure, even under high oscillatory strain (Fig. [137]6E). Fig. 6. [138]Fig. 6 [139]Open in a new tab Construction of Ca-CDs@ALG hydrogel and its therapeutic effect on periodontitis. (A) Schematic illustration of the construction of Ca-CDs@ALG hydrogel. Photographs demonstrating the (B) injectability and (C) plasticity of Ca-CDs@ALG hydrogel taken under sunlight and UV light, respectively. Variations of storage and loss modulus (G’ and G’’, respectively) versus (D) angular frequency and (E) oscillation strain. (F) Images of Ca-CDs@ALG microsphere. Inset: enlarged view of the microspheres with a scale bar of 5 μm. (G) Cumulative release of Ca-CDs from Ca-CDs@ALG microsphere at different time points. (H) Degradation curves of Ca-CDs@ALG microsphere over time. (I) Experimental design for treating rat ligature-induced periodontitis. (J) Representative micro-CT three-dimensional reconstructed images (upper panels) and scanning buccopalatal images (lower panels) along the longitudinal direction of the maxilla. (K) Quantitative analysis of BV/TV, Tb.N, Tb.Th, and Tb.Sp, determined by micro-CT images. Data are presented as mean ± SD from three independent experiments. * indicates P < 0.05, ** indicates P < 0.01, *** indicates P < 0.001. (L) H&E staining showing inflammatory cells infiltration under 4X objective lens (left panels) and a locally enlarged image (right panels). Arrows indicate inflammatory cells. (M) IF staining of IL-1β under 10X objective lens. The white line indicates the boundary of the second molar root. (N) Goldner staining showing alveolar bone height and morphology, depicted by gray lines, under 4X (left panels) and 10X (right panels) objective lens, respectively Nevertheless, inflammatory and immune diseases often result in irregularly shaped defects, which require a delivery system with improved flow properties and small particle sizes. To address this need, a Ca-CDs-based alginate hydrogel microsphere (Ca-CDs@ALG microsphere) system was developed. CLSM images revealed that the microspheres were spherical with an average diameter of approximately 5.7 μm (Fig. [140]6F and S28). The microsphere system facilitated the continuous release of Ca-CDs, with a cumulative release of 75.1% by day 7 (Fig. [141]6G). Additionally, the microspheres degraded by approximately 73.5% after 7 days (Fig. [142]6H). These findings suggest that Ca-CDs can initiate self-crosslinking with alginate to form an injectable, moldable hydrogel for sustained drug release, enhancing the long-term biological effects of Ca-CDs at inflammatory sites and demonstrating significant potential for in vivo applications. Therapeutic efficacy of Ca-CDs and Ca-CDs-based hydrogel in treating periodontitis To comprehensively verify the anti-inflammatory and osteogenic effects of Ca-CDs, and to assess the sustained release properties and long-term biological effects of the Ca-CDs@ALG delivery system, it is essential to select a disease model that closely aligns with these research objectives. Periodontitis, one of the most prevalent chronic inflammatory diseases worldwide, affects the periodontium and is characterized by extensive inflammatory cell infiltration and irregular destruction of the alveolar bone [[143]35, [144]36]. Effective treatment of periodontitis requires drugs that can be precisely injected and fully penetrate the irregular defect areas of periodontal pockets, while also withstanding the flushing effect of saliva to ensure prolonged action and enhanced efficacy. Therefore, a rat periodontitis model was established to fulfill these experimental requirements (Fig. [145]6I and S29). Micro-computed tomography (micro-CT) scans revealed significant alveolar bone height reduction in the periodontitis group, with vertical bone resorption reaching up to one-third of the root, along with horizontal bone loss, confirming the successful establishment of the periodontitis model. In contrast, both the Ca-CDs and Ca-CDs@ALG microsphere groups showed minimal bone resorption, comparable to the control group (Fig. [146]6J). To quantify the extent of bone loss, the distance from the cemento-enamel junction (CEJ) to the alveolar bone crest (ABC) was measured. As shown in Fig. S30, the CEJ-ABC distance in the periodontitis group was 0.77 ± 0.08 mm, while it was significantly reduced by 46.8% in the Ca-CDs group (0.41 ± 0.02 mm) and by 68.9% in the Ca-CDs@ALG microsphere group (0.24 ± 0.03 mm). Further analysis of alveolar bone volume and mineralization levels demonstrated significant increases in bone volume fraction (BV/TV), trabecular number (Tb.N), and trabecular thickness (Tb.Th), along with a notable reduction in trabecular separation (Tb.Sp) in both the treatment groups compared to the periodontitis group (Fig. [147]6K). These findings indicate that treatment with Ca-CDs and Ca-CDs@ALG microsphere can effectively mitigate alveolar bone resorption in inflammatory conditions, with the Ca-CDs@ALG microsphere formulation showing superior therapeutic effects due to its sustained release properties. Histological staining of maxillary specimens was conducted to thoroughly assess the periodontal tissues. Hematoxylin and eosin (H&E) staining in the periodontitis group revealed characteristic pathological changes, including the extension of the junctional epithelium toward the root, detachment of the epithelial attachment from the root surface, and dense infiltration of inflammatory cells (Fig. [148]6L). In contrast, both the Ca-CDs and Ca-CDs@ALG microsphere groups exhibited significantly reduced inflammatory cell infiltration around the epithelium, indicating a reversal of the pathological changes associated with periodontitis. IF staining further demonstrated that treatment markedly reduced IL-1β expression, particularly in the Ca-CDs@ALG microsphere group, which significantly alleviated local inflammation (Fig. [149]6M and S31). Goldner staining showed substantial bone crest resorption close to the root apex and discontinuous trabeculae in the periodontitis group (Fig. [150]6N). In comparison, both the Ca-CDs and Ca-CDs@ALG microsphere groups displayed minimal bone resorption, continuous bone morphology, and significantly upregulated OCN expression, as confirmed by IF staining (Fig. S32). Overall, our therapeutic strategy, which inhibits inflammation and promotes tissue regeneration, presents great potential for the treatment of chronic inflammatory diseases. Additionally, the H&E staining of vital organs in both the Ca-CDs and Ca-CDs@ALG microsphere groups showed no significant differences compared to the control group (Fig. S33), confirming the safety of Ca-CDs and their microsphere formulations for in vivo applications. Discussion The JAK2/STAT3 pathway is recognized as the central communication hub in cytokine signaling, playing a critical role in maintaining immune balance and tissue homeostasis, and has thus emerged as an essential target for drug development [[151]37–[152]40]. Recent clinical trials have shown that targeting JAK2 and preventing its phosphorylation can effectively suppress abnormal immune and inflammatory responses mediated by cytokines, providing solid evidence for the use of JAK2 inhibitors in treating inflammatory and autoimmune diseases [[153]41–[154]43]. Most of the reported research on JAK2 inhibitors has focused on the design and synthesis of small molecules with specific structural characteristics, such as NH, O or CH[2] spacers, heterocyclic moiety, and amide bonds [[155]44–[156]46]. However, these inhibitors usually present limitations, including poor water solubility, low therapeutic concentrations at target sites, and inadequate management of ongoing inflammation-induced bone loss. In response, this study integrates nanomaterial properties with inhibitory functions [[157]47–[158]49], developing a multifunctional JAK2 nanoinhibitor, Ca-CDs, which holds promise as an effective therapeutic strategy. Pharmacophore models and fragment-based drug design can serve as preliminary screening tools to effectively identify key functional groups closely associated with target binding, thereby theoretically excluding inactive structures. This approach not only reduces the workload and cost of experimental validation but also holds significant potential for the design and development of targeted therapeutics. Based on the binding site of JAK2 protein and the structures of existing JAK2 inhibitors, we generated a target-specific training dataset, thereby constructing pharmacophore models and conducting fragment-based drug design. Computational biology confirmed that nitrogen and oxygen-containing aromatic rings facilitate stable binding to JAK2. By leveraging the unique characteristics of CDs, such as their ability to incorporate various functional groups [[159]50–[160]53], and by selecting precursor materials with relevant chemical structures [[161]54–[162]56], we successfully designed and synthesized Ca-CDs, a multifunctional JAK2 nanoinhibitor featuring a highly conjugated sp^2 domain and heteroatom-containing rings [[163]57]. A series of characterization methods, including RNA sequencing, molecular dynamics simulations, and SPR, confirmed that Ca-CDs can stably bind to JAK2, inhibiting JAK2/STAT3 pathway, interrupting cytokine-mediated positive feedback loops, and effectively attenuating excessive inflammatory responses [[164]58]. These findings demonstrate that Ca-CDs possess potent anti-inflammatory properties similar to those of traditional JAK2 inhibitors, but with enhanced biocompatibility and water solubility due to their unique carbon-based nanomaterials [[165]59–[166]62], suggesting that Ca-CDs have the potential to address the challenges faced by conventional JAK2 inhibitors, offering a new approach to managing JAK2-related diseases. A further advantage of Ca-CDs over existing JAK2 inhibitors is their multifunctionality. Traditional JAK2 inhibitors primarily target the JAK2/STAT3 pathway to suppress inflammation but often fail to address the ongoing inflammation-induced bone destruction. In contrast, Ca-CDs demonstrate the capacity to promote bone regeneration under inflammatory conditions through two mechanisms. First, Ca-CDs inhibit inflammatory signaling, mitigating bone loss associated with inflammation. Second, the inherent calcium in Ca-CDs plays a crucial role in the bone matrix, facilitating effective bone regeneration [[167]63]. Furthermore, this calcium can serve as crosslinking junctions, triggering the formation of Ca-CDs@ALG hydrogel. Although alginate hydrogels generally exhibit relatively low mechanical strength, their lower mechanical rigidity actually facilitates complete filling of complex anatomical structure of periodontal pockets, ensuring more uniform drug distribution and localized therapeutic effects. This multifunctionality distinguishes Ca-CDs from existing inhibitors, presenting them as a viable option for inflammatory and immune-related diseases by offering a more comprehensive therapeutic approach. By demonstrating the efficacy of Ca-CDs in reducing inflammation and promoting bone regeneration in a periodontitis model, this study highlights their in vivo potential. Future research should focus on optimizing the design and synthesis of Ca-CDs to improve binding affinity and multifunctionality [[168]64]. Additionally, exploring the potential of Ca-CDs for treating other JAK2-related inflammatory diseases, such as rheumatoid arthritis and inflammatory bowel disease, could expand their therapeutic applications. Conclusion In summary, through the construction of pharmacophore models and fragment-based drug design, we successfully designed and synthesized Ca-CDs as a multifunctional nanoinhibitor targeting JAK2. Strong characterization evidence confirms that Ca-CDs can directly bind to the JAK2, effectively modulating the JAK2/STAT3 pathway and reducing the secretion of pro-inflammatory cytokines. Importantly, the inherent calcium within the Ca-CDs promotes osteogenic differentiation under inflammatory conditions, demonstrating dual anti-inflammatory and osteogenic properties. Additionally, our Ca-CDs serve as crosslinking junctions that trigger the formation of a Ca-CDs-based alginate hydrogel, facilitating long-term retention at local lesion sites and sustained drug release. In vivo studies using a periodontitis model demonstrated that the Ca-CDs and Ca-CDs@ALG delivery system effectively reduced inflammatory cell infiltration and facilitated the repair of inflammation-associated bone defects. Overall, the computer-aided design of Ca-CDs as a JAK2 nanoinhibitor, along with the self-triggering hydrogel delivery system, holds significant potential for treating periodontitis and other inflammatory diseases associated with the JAK2/STAT3 pathway, offering a promising therapeutic strategy for clinical application. Materials and methods Materials Acetylsalicylic acid (99%), metformin (97%) and calcium chloride (99.9%) were purchased from Aladdin (Shanghai, China). Dialysis bag (MWCO = 100–500 Da) was obtained from Spectrum Laboratories (Rancho Dominguez, CA, USA). Dulbecco’s modified Eagle medium (DMEM) with high glucose, alpha minimum essential medium (α-MEM) and penicillin-streptomycin were purchased from Hyclone (Logan, UT, USA). Fetal bovine serum (FBS) was obtained from Clark Bioscience (VA, USA). Cell counting kit-8 (CCK-8) was purchased from Invigentech (Irvine, CA, USA). Cell cycle kit, Hifair^® III 1 st Strand cDNA Synthesis SuperMix for qPCR, Brefeldin A, and Rhod-2 AM were purchased from Yeasen Biotechnology (Shanghai, China), respectively. Annexin V-FITC apoptosis analysis kit obtained from Simu Biotech (Tianjin, China). Lipopolysaccharide (LPS), ascorbic acid, and β-glycerophosphate disodium salt hydrate were purchased from Sigma Aldrich (St Louis, MO, USA). Calcein/PI cell viability/cytotoxicity assay kit, alkaline phosphatase (ALP) assay kit, alizarin red S (ARS) staining solution, modified oil red O staining kit, and enhanced bicinchoninic acid (BCA) protein assay kit were obtained from Beyotime (Shanghai, China). RNAiso Plus reagent was obtained from Takara Bio (Otsu, Japan). Fedratinib, Colivelin and Coumermycin A1 were obtained from MedChenExpress (Monmouth Junction, NJ, USA). Generation of ligand-based common feature pharmacophore models The two-dimensional (2D) structures of 53 known JAK2 inhibitors were obtained from the MedChenExpress database (Fig. [169]S1). Their corresponding three-dimensional (3D) structures were constructed using Avogadro software, followed by optimization through energy minimization within the CHARMM force field. The maximum number of conformations for each molecule was set to 255, and conformations were generated under an energy threshold of 20 kcal/mol. For a molecule with stereo isomers, this threshold applies to conformations for each individual isomer. Subsequently, pharmacophore models were generated using the HipHop algorithm [[170]65]. Ten pharmacophore models were created, and the top two were selected based on their rank values. Fragment-based drug design First, a fragment library was generated from the 53 known JAK2 inhibitors obtained from the MedChenExpress database using the Retrosynthetic Combinatorial Analysis Procedure (RECAP) algorithm [[171]66]. During the library construction, bond-type rules, including amide, ester, amine, urea, ether, olefin, quaternary nitrogen, aromatic nitrogen bonded to aliphatic carbon, lactam nitrogen bonded to aliphatic carbon, aromatic carbon bonded to aromatic carbon, and sulfonamide, were employed to cleave the molecules. Then, the structure of the JAK2 protein, predicted by AlphaFold, was downloaded from the UniProt database (ID: [172]Q62120). After defining the protein’s binding site, the Ludi algorithm was utilized to search the generated fragment library for candidates that bind to the active site. Consequently, suitable fragments capable of interacting with the JAK2 protein were screened, and their positions within the active site of the JAK2 protein structures were visualized using ChimeraX. Synthesis of Ca-CDs A mixture of metformin (1 g) and CaCl[2] (2 g) was dissolved in deionized water (3 mL). This solution was then transferred to a round-bottom flask containing acetylsalicylic acid (0.5 g) and heated in an oven at 120 °C for 110 min. Subsequently, the mixture was reacted at 250 °C under vacuum for 120 min, yielding a yellow, fluffy, porous Ca-CDs sphere. This sphere was dissolved in deionized water (100 mL), sonicated for 5 min to accelerate dissolution, and then centrifuged at 10,000 rpm at room temperature for 10 min to remove large particulate impurities. The supernatant was rotary evaporated and then transferred to a dialysis bag (100–500 Da) for 12 h of dialysis. The purified light yellow Ca-CDs solution was concentrated and freeze-dried to obtain a Ca-CDs solid powder. Undoped CDs were prepared using a similar procedure, except without the addition of CaCl[2]. Following the same approach, CDs synthesized using other organic small molecules as precursors were also prepared. Characterization of Ca-CDs Ultraviolet-visible (UV-vis) absorption spectra were obtained using a Lambda 800 spectrophotometer. Fluorescence and photoluminescence spectra were analyzed with a Shimadzu RF6000 spectrophotometer. High-resolution transmission electron microscopy (HRTEM) image was captured using a JEM-2100 F microscope at an acceleration voltage of 200 kV. Fourier transform infrared (FTIR) spectra were recorded using a Bruker VERTEX 80 V spectrometer. X-ray photoelectron spectroscopy (XPS) was conducted using the NEXSA system from Thermo Scientific. Energy-dispersive spectroscopy (EDS) analysis was conducted using an Oxford Inca X-Max spectrometer. Preparation of Ca-CDs-based alginate hydrogels and microspheres A Ca-CDs solution was added dropwise to sodium alginate (ALG), followed by continuous stirring at 1500 rpm to produce injectable hydrogels. The rheological properties of the hydrogels were analyzed using a DHR-2 rheometer. The variation in storage modulus (G’) and loss modulus (G’’) was monitored across an angular frequency range from 0.1 to 100 rad/s in oscillation mode. Using a similar method, Ca-CDs solution (1 mL) was added dropwise to 20 mL of an aqueous ALG solution at a flow rate of 0.1 mL/min with a syringe pump, followed by magnetic stirring at 400 rpm for 30 min to prepare hydrogel microspheres. The microspheres were alternately washed with deionized water and anhydrous ethanol twice. The size and morphology of the microspheres were characterized using confocal laser scanning microscopy (CLSM). Degradation and release performance of Ca-CDs-based alginate hydrogel microspheres For degradation studies, microspheres were immersed in deionized water at 37 °C. The microspheres were centrifuged and washed at predetermined intervals, followed by freeze-drying and weighing. The degradation rate (D%) of the microspheres was calculated using the formula: D% = (N0 - Nt)/N0 × 100%, where N0 (mg) is the initial weight of the microspheres, and Nt (mg) is their weight after soaking for t days. For release studies, the microspheres were incubated at 37 °C, and 3 mL of supernatant was collected at various time points, while maintaining the total volume at 10 mL by replenishing with deionized water. Given that the concentration of Ca-CDs correlates with their UV-vis absorbance values, the concentration of Ca-CDs released at various time points was determined by measuring the absorbance of the supernatant. The cumulative release rate (%) of Ca-CDs at different time points was calculated as: (Accumulated release of Ca-CDs at the respective time point/Total amount of Ca-CDs) × 100%. Culturing and identification of RAW264.7 cells and rat bone marrow stromal cells (rBMSCs) RAW264.7 cells were resuspended in DMEM high glucose medium supplemented with 10% FBS and cultured in an incubator at 37 °C with 5% CO[2]. Cell density was monitored, and upon reaching 70%−80% confluency, cells were subcultured for further experiments. Male Wistar rats (age 3–4 weeks) were used for isolating rBMSCs employing the whole bone marrow adherence method. The cell suspension was cultured in a 5% CO[2] incubator at 37℃. After 4 days, the medium was replaced with fresh α-MEM complete medium containing 10% FBS, penicillin (100 U/mL), and streptomycin (100 µg/mL). At the third passage, adipogenic and osteogenic differentiation capabilities were assessed using oil red O and ARS staining, respectively. Flow cytometry was employed for phenotypic identification of rBMSCs using CD29 and CD90 as positive surface markers and CD45 and CD34 as negative surface markers. CCK-8 assay for cell toxicity RAW264.7 cells were seeded in a 96-well plate at a density of 5 × 10^3 cells/well. A control group and five experimental groups were established, treating cells with 100 µL of complete medium containing varying concentrations of Ca-CDs (0, 50, 100, 200, 300, and 400 µg/mL). After 24 h, 100 µL of medium containing 10 µL of CCK-8 solution was added to each well. The absorbance at 450 nm was measured using a microplate reader 30 min later, and cell viability was calculated relative to the control group. The same experimental setup and groups were used to assess the viability of rBMSCs seeded at densities of 5 × 10^3 and 2 × 10^3 cells/well, after incubation periods of 24 h and 7 days, respectively. Detection of cell apoptosis and cell cycle by flow cytometry RAW264.7 cells and rBMSCs were seeded in 6-well plates at a density of 2 × 10^5 cells/well and cultured overnight for adherence. The medium was then replaced with complete medium containing different final concentrations of Ca-CDs (0, 100, 200, and 300 µg/mL). After 24 h, cells were collected, washed twice with phosphate buffer saline (PBS), and resuspended in 100 µL of binding buffer. Each sample was stained with 5 µL of Annexin V-FITC and 5 µL of 7-AAD. The samples were incubated for 15 min in the dark, followed by flow cytometry analysis to detect fluorescence signals. rBMSCs were cultured and collected as described above. Cells were stained with PI/RNase staining solution for cell cycle analysis. Live/Dead cell staining RAW264.7 cells and rBMSCs were seeded in confocal dishes at a density of 10^5 cells/well and cultured overnight for adherence. Post-adherence, the medium was replaced with complete medium containing different concentrations of Ca-CDs (0, 100, 200, and 300 µg/mL) and incubated for 24 h. The medium was then discarded, and the cells were stained with 500 µL/well of Calcein/PI staining solution. The cells were incubated at 37 °C in the dark for 30 min before capturing images using CLSM. Cellular uptake of Ca-CDs RAW264.7 cells and rBMSCs were seeded in confocal dishes at a density of 10^5 cells/well and cultured overnight for adherence. After adherence, cells were incubated for 12 h in complete medium containing Ca-CDs at a final concentration of 300 µg/mL. Following incubation, cells were washed three times with PBS to remove any uninternalized material. The internalization of Ca-CDs within the cells was observed under CLSM by examining the fluorescence of the Ca-CDs. Establishment of RAW264.7 cell inflammation model Cells were seeded in culture dishes and divided into seven groups, each with three replicates: (1) control; (2) LPS group; (3) LPS + 100 µg/mL Ca-CDs; (4) LPS + 200 µg/mL Ca-CDs; (5) LPS + 300 µg/mL Ca-CDs; (6) LPS + 300 µg/mL acetylsalicylic acid and metformin (300 µg/mL ASA + MET); (7) LPS + Fedratinib. After cell adherence, the control group’s medium was replaced with complete medium, while the LPS model and treatment groups received complete medium containing LPS (100 ng/mL) to establish the inflammation cell model. Osteogenic induction and model establishment in rBMSCs rBMSCs were seeded and cultured until 70–80% confluency. The medium was then switched to osteogenic induction medium (containing 10 mM β-glycerophosphate, 50 µg/mL ascorbic acid, in α-MEM complete medium) for osteogenic differentiation induction. Normal state rBMSCs grouping: (1) control; (2) 100 µg/mL Ca-CDs; (3) 200 µg/mL Ca-CDs; (4) 300 µg/mL Ca-CDs; (5) 300 µg/mL ASA + MET; (6) Fedratinib; (7) undoped CDs; (8) CaCl[2]. The osteogenic induction medium was replaced every three days. Inflamed state rBMSCs grouping: (1) LPS group; (2) LPS + 100 µg/mL Ca-CDs; (3) LPS + 200 µg/mL Ca-CDs; (4) LPS + 300 µg/mL Ca-CDs; (5) LPS + 300 µg/mL ASA + MET, with five groups in total, each having three replicates. After cell adherence, the control group was switched to osteogenic induction medium, while the LPS model and treatment groups received complete medium containing LPS (10 µg/mL) to establish the inflammation model. Osteogenic induction medium was replaced every three days. Reverse transcriptase quantitative polymerase chain reaction (RT-qPCR) RAW264.7 cells and rBMSCs were seeded in 6-well plates at a density of 2 × 10^5 cells/well and cultured as per the above methods for 1 day and 7 days, respectively. RNA was isolated using 1 mL RNAiso Plus reagent per well. RNA concentration was determined using the Thermo Scientific NanoDrop spectrophotometer. RNA was reverse transcribed into cDNA using the Hifair^® III 1 st Strand cDNA Synthesis SuperMix for qPCR. RT-qPCR was conducted on a Roche LC96 fluorescent quantitative PCR instrument, analyzing gene expression levels with Actb as the reference gene. The primers are listed in Tables S1 and S2. Western blot RAW264.7 cells were seeded in 6 cm diameter culture dishes at a density of 10^6 cells/dish and grouped according to the inflammation cell model. After 20 h of culture with appropriate treatments, Brefeldin A was added to a final concentration of 1 µg/mL, followed by a further 4-hour incubation. Cells were lysed using RIPA lysis buffer containing protease and phosphatase inhibitors. Protein concentration was determined using the enhanced BCA protein assay kit. Proteins (30 µg) were separated using 10% SDS-PAGE and transferred to polyvinylidene fluoride (PVDF) membranes. Membranes were blocked with 5% BSA for 1 h to prevent nonspecific binding, followed by overnight incubation at 4 °C with primary antibodies: anti-GAPDH (10494-1-AP, 1:5000, Proteintech), anti-IL-1β (ab254360, 1:1000, Abcam), anti-iNOS (ab178945, 1:1000, Abcam), anti-TNF-α (17590-1-AP, 1:1000, Proteintech), anti-β-actin (66009-1-Ig, 1:5000, Proteintech), anti-JAK2 (3230, 1:1000, CST), anti-p-JAK2 (4406, 1:1000, CST), anti-STAT3 (4904, 1:2000, CST), and anti-p-STAT3 (9145, 1:2000, CST). Membranes were washed and incubated with secondary antibodies for 2 h, followed by ECL detection and quantification using ImageJ. rBMSCs were seeded in 9 cm diameter culture dishes at a density of 2 × 10^6 cells/dish and subjected to the same treatment protocol as defined in the established cell model for a duration of 7 days. Total proteins were extracted and incubated with primary antibodies: anti-GAPDH, anti-RUNX2 (ab236639, 1:1000, Abcam), anti-SP7 (ab209484, 1:1000, Abcam), and anti-OCN (23418-1-AP, 1:1000, Proteintech), following the same procedure as above, with quantification using ImageJ. ALP and ARS staining rBMSCs were seeded in 12-well plates at a density of 10^5 cells/well and cultured until 70–80% confluency. Cell models under normal and inflammatory conditions were established as previously described, and cultured for 7, 14, and 21 days, respectively. The cells were then stained using a kit according to the manufacturer’s instructions. The stained cells were observed and imaged using an inverted microscope. Immunofluorescence (IF) staining rBMSCs were seeded at a density of 10^5 cells/well in confocal dishes. The cells were then cultured for 14 days in osteogenic induction medium containing LPS either 0–300 µg/mL Ca-CDs. Following this period, the cells were fixed with 4% paraformaldehyde for 10 min and permeabilized with 0.1% Triton X-100 for 10 min. Nonspecific binding sites were blocked with 5% goat serum for 1 h. Subsequently, cells were incubated overnight at 4 °C with anti-OCN (23418-1-AP, 1:400, Proteintech). The next day, cells were incubated with CoraLite594-conjugated secondary antibody (SA00013-4, 1:200, Proteintech) and Phalloidin-FITC (CA1620, 1:200, Solarbio) in the dark for 1 h. Images were captured using CLSM. Measurement of intracellular Ca^2+ rBMSCs were plated in confocal dishes at a density of 10^5 cells/well. The cells were treated with Ca-CDs or CaCl[2] for 12 h, followed byincubation with Rhod-2 AM (2 µM) for 30 min, and subsequently imaged using CLSM. Transcriptome analysis by RNA sequencing The RAW264.7 inflammation cell model was established with two groups: LPS and LPS + 300 µg/mL Ca-CDs. RNA was extracted after 24 h of culture and subjected to transcriptome sequencing performed by Shanghai OE Biotech Co., Ltd. using the Illumina NovaSeq 6000 platform, generating 150 bp paired-end reads. Approximately 50 M raw reads were obtained for each sample. The raw reads in FASTQ format were processed to remove low-quality reads, obtaining clean reads for subsequent data analysis. The reference genome alignment was performed using HISAT2 software, and gene expression levels were calculated. Reads count for each gene was obtained using HTSeq-count and normalized across samples using DESeq2 software. Differential expression was determined with a threshold of q-value < 0.05 and fold change > 2. Gene Ontology (GO) enrichment and Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway enrichment analysis of differentially expressed genes (DEGs) were conducted using R, based on the hypergeometric distribution. Molecular dynamics simulations The structure of the JAK2 protein, predicted by AlphaFold, was downloaded from the UniProt protein database (ID: [173]Q62120) and subsequently assessed for quality using Ramachandran plots. Based on analyses from HRTEM and XPS, a simplified model of Ca-CDs was constructed and optimized using density functional theory methods. Molecular dynamics simulations were performed using the GROMACS software package version 2018.8 under the AMBER99SB-ILDNP force field. The system was solvated in a cubic box with transferable intermolecular interaction potential 3 points (TIP3P) water molecules, ensuring a minimum distance of 10 Å between the protein and the box edges. Extra system charges were neutralized with Cl- ions. Periodic boundary conditions were applied in all directions to simulate an infinite system. Steepest descent and conjugate gradient minimization were conducted (tolerance up to 1000 kJ/mol/nm) to eliminate any unfavorable interactions. Prior to production simulation, the system underwent two equilibration steps where all atoms, except for water molecules and Cl- ions, were restrained to prevent conformational changes. Initially, the system was simulated under constant volume (NVT ensemble) conditions, reaching 300 K within 1 ns using the V-rescale method. Subsequently, the equilibrated structure of the NVT ensemble was balanced under constant pressure (NPT ensemble) at isotropic pressure of 1.0 bar using the Parrinello-Rahman barostat for 1 ns. The production Molecular dynamics simulation was carried out for 120 ns without any restraints. Simulation trajectories were saved every 10 ps for analysis of root mean square deviation (RMSD) and root mean square fluctuation (RMSF). Surface plasmon resonance (SPR) SPR assays were conducted to evaluate the interaction between purified recombinant JAK2 proteins and Ca-CDs using a Biacore 8 K instrument (Cytiva). The His-tagged human JAK2 protein was expressed in E. coli (HY-P7989, MedChenExpress) and exhibited 96.3% homology with the mouse JAK2 protein. This recombinant His-JAK2 was purified and immobilized on a CM5 chip under optimized conditions. The running buffer was prepared with the following composition: 50 mM Tris-HCl, 150 mM NaCl, 5 mM MgCl[2], and 0.05% Tween-20. The Ca-CDs were serially diluted in running buffer, covering a range of concentrations, and injected into the sensor chip. Each binding cycle consisted of a 60-second binding phase followed by a 60-second dissociation phase, with a flow rate of 30 µL/min. The equilibrium dissociation constant (K[D]) was calculated using the Biacore 8 K Evaluation Software, employing appropriate fitting models to ensure accuracy. Establishment of rat periodontitis model Twenty-four male Wistar rats, weighing 150–200 g, were randomly divided into four groups: control, periodontitis, Ca-CDs, and Ca-CDs@ALG microsphere. Periodontitis was induced in the second upper molar using the ligature method. Both the control and periodontitis groups received 20 µL of PBS injected into the gingival sulcus, while the Ca-CDs group received 20 µL of Ca-CDs solution (300 µg/mL), and the Ca-CDs@ALG microsphere group received 20 µL of microsphere (300 µg/mL). Treatments were administered every three days. After 28 days, the rats were anesthetized, perfused with saline and 4% paraformaldehyde, and samples of the maxilla, heart, liver, spleen, and kidneys were collected. All animal experiments were conducted following Jilin University’s committee guidelines for animal experiments (approval number: 20240290). Micro-computed tomography (Micro-CT) bone tissue analysis The collected maxillary specimens were scanned using µCT-50 (Scanco Medical AG, Bassersdorf, Switzerland) with scanning parameters of 70 kVp and 200 µA. ImageJ software was used to measure the distance between the cemento-enamel junction (CEJ) and the alveolar bone crest (ABC) at four points on the buccal and palatal sides of the second upper molar. Bone tissue parameters were also obtained for each group. Histological staining analysis Maxillary and visceral specimens, fixed and decalcified in 4% paraformaldehyde, were dehydrated in an ethanol gradient, embedded in paraffin, and sectioned. Sections were stained according to the kit instructions for Hematoxylin and eosin (H&E), Goldner’s trichrome, and IF antibodies included anti-OCN (23418-1-AP, 1:400, Proteintech) and anti-IL-1β (ab254360, 1:50, Abcam). Statistical analysis Each experiment was conducted at least three times. Data are presented as mean ± standard deviation. All statistical analyses were performed using Origin software. For comparisons involving multiple groups, statistical analysis was performed using one-way ANOVA, with a P-value of < 0.05 considered to indicate statistical significance. Supplementary Information [174]Supplementary Material 1^ (28.2MB, docx) [175]Supplementary Material 2^ (8.7MB, mp4) [176]Supplementary Material 3^ (8.3MB, mp4) [177]Supplementary Material 4^ (12MB, mp4) [178]Supplementary Material 5^ (12.5MB, mp4) [179]Supplementary Material 6^ (9.5MB, mp4) Author contributions Yang Liu: Writing - review & editing, Writing - original draft, Validation, Supervision, Software, Methodology, Investigation, Formal analysis, Data curation, Conceptualization. Duan Wang: Visualization, Supervision, Methodology, Investigation, Data curation. Wenqian Zheng: Visualization, Methodology, Formal analysis. Yuxuan Wang: Software, Validation, Investigation. Peiyu Wang: Visualization, Methodology, Investigation. Yuping Zhao: Visualization, Methodology. Xiaofan Liu: Visualization, Methodology. Yaru Shi: Investigation, Formal analysis. Yi Wang: Writing - review & editing, Supervision. Na Zhou: Writing - review & editing, Investigation. Fermín E. González: Writing - review & editing, Supervision. Hongchen Sun: Writing - review & editing, Resources. Ding Zhou: Writing - review & editing, Writing - original draft, Supervision, Software, Resources, Project administration, Funding acquisition, Conceptualization. Xiaowei Xu: Writing - review & editing, Writing - original draft, Validation, Resources, Project administration, Funding acquisition. All authors have read and approved the final manuscript. Funding This study was supported by grants from the National Key Research and Development Program for Young Scientists (2022YFA1105800), the National Natural Science Foundation of China (82470957), Jilin Province Science and Technology Research (Project No. YDZJ202501ZYTS020), and the Bethune Plan Project of Jilin University (2024B14). Data availability No datasets were generated or analysed during the current study. Declarations Ethics approval and consent to participate All animal experiments were conducted following Jilin University’s committee guidelines for animal experiments (approval number: 20240290). Competing interests The authors declare no competing interests. Footnotes Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. Yang Liu and Duan Wang contributed equally to this work. Contributor Information Ding Zhou, Email: zhouding@jlu.edu.cn. Xiaowei Xu, Email: xiaoweixu@jlu.edu.cn. References