Abstract With the threat posed by drug‐resistant pathogenic bacteria, developing non‐antibiotic strategies for eradicating clinically prevalent superbugs remains challenging. Ferroptosis is a newly discovered form of regulated cell death that can overcome drug resistance. Emerging evidence shows the potential of triggering ferroptosis‐like for antibacterial therapy, but the direct delivery of iron species is inefficient and may cause detrimental effects. Herein, an effective strategy to induce bacterial nonferrous ferroptosis‐like by coordinating single‐atom metal sites (e.g., Ir and Ru) into the sp^2‐carbon‐linked covalent organic framework (sp^2c‐COF‐Ir‐ppy[2] and sp^2c‐COF‐Ru‐bpy[2]) is reported. Upon activating by light irradiation or hydrogen peroxide, the as‐constructed Ir and Ru single‐atom catalysts (SACs) can significantly expedite intracellular reactive oxygen species burst, enhance glutathione depletion‐related glutathione peroxidase 4 deactivation, and disturb the nitrogen and respiratory metabolisms, leading to lipid peroxidation‐driven ferroptotic damage. Both SAC inducers show potent antibacterial activity against Gram‐positive bacteria, Gram‐negative bacteria, clinically isolated methicillin‐resistant Staphylococcus aureus (MRSA), and biofilms, as well as excellent biocompatibility and strong therapeutic and preventive potential in MRSA‐infected wounds and abscesses. This delicate nonferrous ferroptosis‐like strategy may open up new insights into the therapy of drug‐resistant pathogen infection. Keywords: anti‐infection therapy, covalent organic frameworks, ferroptosis‐like, single‐atom catalysts, transition metal complex __________________________________________________________________ A novel nonferrous ferroptosis‐like strategy by coordinating single‐atom metal sites (e.g., Ir and Ru) into the sp^2‐carbon‐linked covalent organic framework is reported. The as‐constructed Ir and Ru sites can effectively induce lipid peroxidation‐driven bacterial ferroptotic damage, showing strong therapeutic and preventive potential in infected wounds and abscesses. graphic file with name ADVS-10-2207507-g005.jpg 1. Introduction Antimicrobial resistance severely compromises traditional chemotherapy regimens and persists as a worldwide health problem.^[ [42]^1 ^] Bacteria can employ a variety of strategies to avert growth suppression of conventional antibiotic therapy, consisting of enzyme inactivation and target modification.^[ [43]^2 ^] Furthermore, they could spark post‐antibiotic dilation and recurrent/persistent infection through reprogramming host metabolism, interfering with degradation pathways, and inhibiting immune cells.^[ [44]^3 ^] Alternative strategies capable of circumventing antibiotic resistance are of utmost importance. Ferroptosis is an iron‐dependent form of cell death resulting from lipid peroxidation (LPO) that has been implicated in various biological contexts, from development to aging to immunity and cancer.^[ [45]^4 ^] Several types of strategies have been described to induce ferroptosis to date, comprising iron delivery,^[ [46]^5 ^] system Xc^− suppression,^[ [47]^6 ^] glutathione (GSH) depletion, and glutathione peroxidase 4 (GPX4) inhibition.^[ [48]^7 ^] The precise mechanism that eventually results in ferroptotic cell death probably involves damage to membrane integrity, disruption of membrane properties through lipid cross‐linking, and further oxidative impairment to macromolecules and cellular structures induced by reactive oxygen species (ROS) derived from polyunsaturated fatty acid chains.^[ [49]^8 ^] It has been hypothesized that the regulation of key molecules in classical regulatory pathways of ferroptosis can serve as a potential approach to overcoming drug resistance.^[ [50]^9 ^] Emerging ferrous‐based nanomaterials, comprising ferumoxytol, nano‐iron sulfides, and iron–organic frameworks, have been applied as inducers for ferroptosis benefiting from the Fenton reaction impelled by Fe^2+.^[ [51]^10 ^] For instance, attempts have been made to introduce iron ions into the metastable Fe[3]S[4] (greigite) or FeSO[4] to facilitate the iron overload‐triggered GSH consumption, leading to the ferroptosis‐like of bacterial cells.^[ [52]^11 ^] Ferroferric oxide‐based nanoassemblies were also utilized as the inducers for bacterial ferroptosis‐like death by eliciting intracellular iron overload and iron metabolism interference.^[ [53]^12 ^] However, these current iron‐based nanomaterials used to induce bacterial ferroptosis‐like are far from satisfactory, requiring very high Fe doses or supplementary ingredients to achieve combinatorial effects in general.^[ [54]^10b ^] Alternatively, the direct delivery of iron species may cause detrimental effects such as neurovirulence, oxidative stress, and anaphylactic reactions in normal tissues.^[ [55]^13 ^] Single‐atom catalysts (SACs) have become an exciting frontier recently in chemical catalysis because of their precisely identified active centers, robust catalytic performances, and high stability.^[ [56]^14 ^] SACs can be regarded as the extreme limit of the precise design of nanocatalytic materials at the atomic level. Particularly, they have been utilized as bio‐inspired nanozymes to mimic natural enzymes’ structure and excellent catalytic ability, efficiently generating excessive ROS for bacterial or tumor inhibition.^[ [57]^15 ^] For example, Qu and colleagues report a self‐adapting iron‐based SAC to accelerate selective and safe ferroptosis;^[ [58]^16 ^] a nonferrous‐based Pd‐SAC with simulated activities of double peroxidase (POD) and glutathione oxidase (GSHOx) also efficiently induces ferroptosis characterized by upregulation of LPO and ROS.^[ [59]^17 ^] Unfortunately, considering the low intracellular H[2]O[2] level of bacterial cells, Fenton reaction alone is difficult to generate sufficient ROS, which weakens the catalytic therapeutic efficacy of conventional SACs.^[ [60]^18 ^] Recently, isolated active metal centers anchored to solid support represent an innovative breakthrough in photochemistry.^[ [61]^19 ^] Nanoscale covalent organic frameworks (COFs) composed of suitable building blocks and organic functional groups have emerged as highly promising carriers for their tunable microstructures and optically electrical properties in preference to traditional catalyst supports.^[ [62]^20 ^] The possibility of utilizing COFs as supporting materials to construct SACs to meet the requirements of ferroptosis was investigated. Various monatomic metal centers anchored on COFs could present effective photocatalysis.^[ [63]^21 ^] Many studies have certified that transition‐metal elements such as Ir and Ru could act as single‐atom active sites without disrupting the framework for constructing high‐performance photochemical catalysts.^[ [64]^22 ^] Adding single transition metal atoms to the bare photocatalysts can expand the optical response range, shorten the electron transfer distance, and form stable intermediate configurations in photocatalytic reactions by virtue of the increased delocalization effect, enduing SACs with excellent photocatalytic performance.^[ [65]^21 , [66]^23 ^] Transition metal SACs have also been reported to exhibit POD and GSHOx activity.^[ [67]^24 ^] Hence, exploring a COF‐based SAC paradigm for bacterial ferroptosis‐inducing agents is imperative and highly desirable. Herein, we prepare two types of monatomic transition metal sites (e.g., Ir and Ru) anchored on sp^2c‐linked COF (sp^2c‐COF) skeletons with a metal–nitrogen–carbon bridging structure (Scheme [68] 1 ). Facilitated by covalent interactions in the Schiff base reaction, methoxy polyethylene glycol amine (mPEG‐NH[2]‐4000) polymer could be coated to produce the hydrophilic and highly biocompatible SACs (sp^2c‐COF‐Ir‐ppy[2] and sp^2c‐COF‐Ru‐bpy[2]). The experimental results and density functional theory (DFT) calculation indicated that the excellent photocatalytic capacity and POD activity of Ir and Ru SACs were attributed to the intrinsic porous properties of the COF and the synergistic effect between atomically dispersed metal centers and sp^2c‐COF hosts. Upon irradiation, the Ir and Ru active sites could cause the production of suprathreshold ROS, the consumption of intracellular GSH, and the disturbance of respiratory chain and metabolism together facilitating irreversible LPO‐driven ferroptosis‐like pathways. Both inducers show low hemolysis and cytotoxicity, as well as potent antibacterial activity against various bacteria, drug‐resistance bacteria, and strong therapeutic and preventive potential for methicillin‐resistant Staphylococcus aureus (MRSA)‐induced infections in the wound and abscess models. Collectively, we conducted a proof‐of‐concept study to discover COF‐based SAC as an antibacterial ferroptosis‐like initiator to eliminate infections. Scheme 1. Scheme 1 [69]Open in a new tab Scheme depicting coordinative synthesis of Ir and Ru SACs in the pore of sp^2c‐COF. 2. Results and Discussion 2.1. Synthesis and Characterization Substantial attention has recently been focused on the olefin‐based COFs with fully π‐conjugated systems, a new class of promising semiconductor materials.^[ [70]^25 ^] The robust C=C bond not only endues the framework with excellent stability under harsh conditions but also ensures efficient electron transfer via extensive π‐conjugation throughout the framework.^[ [71]^26 ^] Nevertheless, due to the poor reversibility of C=C bond formation, the preparation of sp^2‐carbon conjugated COFs maintains a tremendous challenge. Fortunately, under solvothermal conditions (tetrahydrofuran/0.1 m Cs[2]CO[3] = 20/1 v/v, 3 days, 120 °C), the topology‐directed polycondensation of C [3]‐symmetric 4,4′,4″‐(1,3,5‐triazine‐2,4,6‐triyl)tribenzaldehyde (TA) as a knote and C [2]‐symmetric linear 2,6‐dicyanomethylbenzo[1,2‐d:4,5‐d′]bisthiazole (BTHAN) as a linker yielded a sp^2c‐COF (Figures [72]S1–S4, Supporting Information). The establishment of a donor–acceptor (D–A) conjugated system in COF by covalent linking of sulfur‐containing aromatic heterocyclic sites with triazine active sites would be an effective strategy to enhance intramolecular charge transfer capability to promote π‐electron delocalization, resulting in a narrower bandgap and desirable highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) energy levels.^[ [73]^27 ^] This push–pull effect was demonstrated by calculating the molecular orbital density of the ligand by Gaussian at the B3LYP/6‐31G* theoretical level (Figure [74] 1a). The HOMO orbitals of BTHAN are mainly concentrated on the benzothiazole portion, while the LUMO is delocalized mainly over the triazine units, suggesting a much higher degree of LUMO–HOMO separation, leading to more efficient charge separation. Figure 1. Figure 1 [75]Open in a new tab Characterizations of sp^2c‐COF. a) HOMO and LUMO of BTHAN and TA. b) Extended structures of non‐interpenetrated sp^2c‐COF. c) Electrostatic potential surface of sp^2c‐COF repetitive units showing possible active sites. d) The charge density differences of sp^2c‐COF and its magnified top and side views, with magenta and yellow representing electron accumulation and depletion, respectively. e) Hole and electron distribution heat map in the excited state of sp^2c‐COF fragment. f) The direction of the dipole moment produced by the local polarization of sp^2c‐COF. g) The electron–hole distributions of sp^2c‐COF in the excited state, with blue and green representing electron accumulation and depletion, respectively. h) The Pawley refinement of sp^2c‐COF: the experimental XRD pattern is shown in black, the Pawley refinement pattern in red, their difference in blue, and the simulated pattern using AA stacking mode in green. i) Crystal structures of the sp^2c‐COF AA stacking model. j) N[2] adsorption–desorption isotherms for sp^2c‐COF at 77 K. Inset: pore‐size distribution calculated by fitting the NLDFT model to adsorption data. k) The simulation pore size of sp^2c‐COF in AA stacking. l,m) AFM image and its 3D topography image of sp^2c‐COF. n,o) Surface potential image and its 3D potential lines profile of sp^2c‐COF. For the observation and comparison of local polarization and charge separation behavior of the synthesized COF molecules, we first exploited DFT to study the optimized molecular structures (Figure [76]1b,c). Noticeably, the electrostatic potential analysis for sp^2c‐COF exhibits a positive distribution mainly on TA units, while the negative regions are mainly located on the cyanide group of the BTHAN portion, indicating that the positive and negative charges were obviously separated due to the effect of polarization on the local charge density triggered by the D–A characteristics of the backbones.^[ [77]^28 ^] Afterward, the charge separation ability of COFs in ground state and excited state was investigated. First, the benzothiazole group was recognized as an electron accumulator of sp^2c‐COF in the ground‐state charge density differences (Figure [78]1d), yet the electrons in the TA group were depleted, indicating an uneven charge distribution. Thereafter, BTHAN can create more holes and electrons according to the heat map for sp^2c‐COF fragments (Figure [79]1e). In addition, the dipole moment of the TA to BTHAN molecule was calculated as 17.98 Debye, demonstrating the presence of local charge polarization (Figure [80]1f). The excited‐state charge‐separation behavior of sp^2c‐COF was analyzed with the electron–hole distribution, in which an obvious spatial charge separation was examined (Figure [81]1g). To sum up, these results demonstrate that the local polarization in a single sp^2c‐COF molecule can lead to significant charge separation. The powder X‐ray diffraction (PXRD) measurement exhibits three distinct peaks at 2.390, 4.633, and 21.817, indexed as the (100), (200), and (001) reflections, respectively (Figure [82]1h, black curve). The optimization for the 2D monolayer conformation and configuration of different stacking models has been conducted via the density function‐based tight binding (DFTB) method. The AA‐stacking model with the most favorable energy was obtained and the yielded PXRD pattern (Figure [83]1h, green curve) is consistent with the profile examined in the experiment. The Pawley‐refined PXRD pattern (Figure [84]1h, red curve) with the space group P6/m and unit cell parameters of a = b = 44.6378 Å, c = 3.5311 Å, and c/a = 0.0791 reproduced the curve observed in the experiment with negligible differences (Figure [85]1h, blue curve). Table [86]S1 of the Supporting Information summarizes atomic atomistic coordinates generated by DFTB calculation and Pawley refinement, respectively. Hence, the reconstructed sp^2c‐COF reveals an extended hexagonal 2D lattice with sp^2 carbon skeleton along the x and y directions (Figure [87]1i). The existence of the (001) plane at 21.817° indicates the structural order of 3.5 Å separation in the z direction perpendicular to the 2D sheet. Fourier transform infrared (FT IR) spectroscopy disclosed that for both BTHAN monomer and sp^2c‐COF, the cyano group exhibits a stretching vibrational peak at 2248 cm^−1 (Figure [88]S5, Supporting Information). A peak at 1704 cm^−1 ascribed to the C=O stretching vibration was observed in monomer TA and was widely attenuated in sp^2c‐COF, suggesting that there was a high degree of polymerization in the skeleton. The newly formed peak at 3047 cm^−1 in sp^2c‐COF can be attributed to CH=C stretching, which clearly indicated the C=C connection in the skeleton. Solid‐state ^13C cross‐polarization magic‐angle spinning nuclear magnetic resonance (^13C CP‐MAS NMR) spectroscopy showed a peak at 163 ppm, confirming the presence of a thiazole ring (Figure [89]S6, Supporting Information). Peaks of ≈105 and 115 ppm further supported the formation of vinylene bonds and the presence of cyano units.^[ [90]^29 ^] These remarkable features manifested the successful condensation of the monomers. Nitrogen adsorption–desorption tests were carried out at 77 K and the characteristic type I curves have been recognized in the simulated shape diagram (Figure [91]1j). At low relative pressure, gas absorption increases sharply (P/P [0] < 0.1), which indicates the presence of micropores. The Brunauer–Emmett–Teller surface area was calculated to be 209.5 m^2 g^−1. The average aperture obtained by DFT fitting is ≈4.2673 nm, which corresponds to the theoretical value (4.8835 nm) in AA models (Figure [92]1k). AFM displayed that the surface roughness of sp^2c‐COF is around tens of nanometers (Figure [93]1l,m). The significant difference in the surface potential of the sp^2c‐COF layer in vertical‐contact mode demonstrates remarkable local polarization‐induced charge separation characteristics (Figure [94]1n,o).^[ [95]^30 ^] Finally, thermogravimetric analysis (TGA) suggested that sp^2c‐COF performs excellent thermal stability, even with a residual carbon content of more than 50% up to 800 °C (Figure [96]S7, Supporting Information). Postsynthetic metalation of sp^2c‐COF involved dispersing it in dichloromethane or dichloromethane/methanol in the presence of dimer [Ir[2](ppy)[4]Cl[2]] and dimer [Ru[2](bpy)[2]Cl[2]] (Figure [97] 2a). For this metalation, dimeric transition complexes were chosen as iridium and ruthenium sources since they are capable of binding powerful photoredox complexes through coordination with the thiazole and cyanogen ligands present in sp^2c‐COF.^[ [98]^31 ^] The content of Ir and Ru incorporated into the metalized sp^2c‐COF (sp^2c‐COF‐Ir‐ppy[2]and sp^2c‐COF‐Ru‐bpy[2]) was analyzed with an inductively coupled plasma optical emission spectrometer (ICP‐OES). The contents of Ir and Ru in COFs were 5.37 and 2.12 wt%, respectively. These values imply that 31.6% of the thiazole and cyan ligands in sp^2c‐COF are coordinated with Ir and 23.7% with Ru. Interestingly, metal absorption in the solution is quite efficient. In reality, 50% of the Ir species in solution are bound to the material, whereas 22% of the Ru precursors are bound to the framework. In addition, FT IR, ^13C CP‐MAS NMR, and TGA for sp^2c‐COF‐Ir‐ppy[2] and sp^2c‐COF‐Ru‐bpy[2] showed no significant differences between the original and metalized COFs (see the Supporting Information). Note that the zeta potential of sp^2c‐COF increased positively from −28.1 to −10.2 and +16.8 mV, after postsynthetic metalation of Ir and Ru (Figure [99]S8, Supporting Information). The stability of sp^2c‐COF‐Ir‐ppy[2] and sp^2c‐COF‐Ru‐bpy[2] was also proven in neutral and weakly acidic physiological mediums, such as PBS (0.1 m, pH 7.4) and PBS (0.1 m, pH 6.0). As demonstrated in Figure [100]S9 of the Supporting Information, both samples demonstrated superior dispersibility and stability in a given medium for over one week without forming any significant aggregations. Furthermore, the metal‐centered atoms in sp^2c‐COF‐Ir‐ppy[2] and sp^2c‐COF‐Ru‐bpy[2] were stable and negligible Ir or Ru‐release was observed for one‐week incubation in both neutral and weakly acidic conditions (Figure [101]S9, Supporting Information). Figure 2. Figure 2 [102]Open in a new tab Morphologies of Ir and Ru SACs. a) Synthetic route of sp^2c‐COF‐Ir‐ppy[2] and sp^2c‐COF‐Ru‐bpy[2]. b–d) FESEM images of b) sp^2c‐COF, c) sp^2c‐COF‐Ir‐ppy[2], d) and sp^2c‐COF‐Ru‐bpy[2]. e–g) TEM and HRTEM images of e) sp^2c‐COF, f) sp^2c‐COF‐Ir‐ppy[2], g) and sp^2c‐COF‐Ru‐bpy[2]. h–j) TEM‐EDX mapping images of C, N, Ir, and Ru elements in selected areas of h) sp^2c‐COF, i) sp^2c‐COF‐Ir‐ppy[2], j) and sp^2c‐COF‐Ru‐bpy[2] (dark field mode). k–m) SAC‐HAADF‐STEM images of k) sp^2c‐COF, l) sp^2c‐COF‐Ir‐ppy[2], and m) sp^2c‐COF‐Ru‐bpy[2]. The field emission scanning electron microscope (FESEM) images exhibited in Figure [103]2b–d manifest the similar layered morphology of three COFs, which are hundreds of nanometers in size. The transmission electron microscope (TEM) image reveals that sp^2c‐COF, sp^2c‐COF‐Ir‐ppy[2], and sp^2c‐COF‐Ru‐bpy[2] possess a ribbon‐like layered structure, in good agreement with FESEM (Figure [104]2e–g). High‐resolution TEM (HRTEM) characterization of the same sample shows a cellular internal structure of COFs, with bright spots corresponding to the pores. The presence of evenly distributed Ir and Ru atoms in the framework was demonstrated using energy‐dispersive X‐ray spectroscopy in scanning transmission electron microscopy (TEM‐EDX) (Figure [105]2h–j). No Ir, Ru (or oxide) nanoparticles or clusters are detected in the spherical aberration‐corrected high‐angle annular dark‐field scanning TEM (SAC‐HAADF‐STEM) image. Single Ir and Ru atoms are identified in the SAC‐HAADF‐STEM image using Z‐contrast (Figure [106]2k–m).^[ [107]^32 ^] Furthermore, X‐ray absorption fine structure spectroscopy (XAFS) tests were conducted to certify the dispersion and coordination environment of single Ir and Ru atoms. Compared with IrO[2]/Ir foil and RuO[2]/Ru‐foil as references, the X‐ray absorption near‐edge structure