Pancatalytic biomaterials enable inflammation-related disease intervention

Xiaoyan Jiang ab and Yu Chen *abc
aSchool of Medicine, Shanghai University, Shanghai, 200444, P. R. China. E-mail: chenyuedu@shu.edu.cn
bMaterdicine Lab, School of Life Sciences, Shanghai University, Shanghai 200444, P. R. China
cShanghai Institute of Materdicine, Shanghai 200051, P. R. China

Received 10th June 2025 , Accepted 20th August 2025

First published on 22nd August 2025


Abstract

Inflammatory disease (ID) is an umbrella term encompassing all conditions where both acute and chronic inflammation serve as the central pathogenic mechanism, imposing a profound global health burden due to the high morbidity, mortality, and socioeconomic impact of these conditions. These pathologies are unified by dysregulated inflammatory cascades, oxidative stress, and immune microenvironment imbalances, which drive progressive tissue damage and organ dysfunction. Conventional therapies, including immunosuppressants and anti-inflammatory agents, often provide only palliative relief, fail to address root causes, and carry the risk of systemic toxicity. The emerging pancatalytic biomaterial based approach, which involves the holistic management of preparation (P) of the catalyst, activation (A) of the biological effect and nontoxic treatment (N) of diseases, offers a transformative paradigm for precision management of ID. These pancatalytic biomaterials constitute the operational foundation of pancatalytic therapy, a multidisciplinary framework that integrates catalytic biomaterials with catalytic biology and catalytic medicine to systematically orchestrate pathological processes via the P–A–N framework, ultimately redefining therapeutic strategies from toxic therapy to nontoxic treatment. This review systematically investigates the rational design of pancatalytic biomaterials, classifying them into inorganic, organic, and organic/inorganic hybrid systems based on the composition. The therapeutic applications of pancatalytic biomaterials across cardiovascular, neurological, respiratory, sensory and other systems are comprehensively summarized. Finally, the current challenges in this field are discussed, and future developments and trends in employing pancatalytic biomaterials for inflammatory disease treatment are projected, thereby advancing further research and clinical translation.


image file: d5nh00408j-p1.tif

Xiaoyan Jiang

Xiaoyan Jiang received her master's degree from Shanghai University. Currently, she is pursuing her PhD degree under the supervision of Prof. Yu Chen at Shanghai University. Her research focuses on developing nanomedicine approaches for sensory system disorders (particularly in visual and auditory systems) and tissue repair applications.

image file: d5nh00408j-p2.tif

Yu Chen

Yu Chen received his PhD degree from the Shanghai Institute of Ceramics, Chinese Academy of Sciences (SICCAS). He is now a full professor at Shanghai University. His research focuses on materdicine, nanomedicine, and nanobiotechnology, which involve the design, fabrication, and biomedical applications of mesoporous nanoparticles, two-dimensional nanosheets, and 3D printing bioscaffolds. The focused biomedical applications include drug/gene delivery, molecular imaging, pancatalytic medicine, immunotherapy, tissue engineering, and in situ localized disease therapy. He has published more than 400 scientific papers in the nanomedicine field with a total citation of more than 53[thin space (1/6-em)]000 times (h-index: 122). We are honored to participate as invited contributors in celebrating the 10th anniversary of Nanoscale Horizons. Our collaboration with the journal began with our first publication in 2019, followed by another significant contribution in 2020. These works collectively established our research foundation in the field of nanocatalytic biomaterials. We sincerely value the prestigious platform that Nanoscale Horizons provides, which has enabled us to share our latest perspectives on pancatalytic biomaterials for inflammatory disease therapy in our current contribution. Throughout our collaboration, the journal has served as a vital conduit for promoting interdisciplinary dialogue between nanotechnology and medicine, while effectively disseminating our scientific achievements. We look forward to maintaining this productive collaboration in the coming years.


1. Introduction

Inflammation-related diseases (IDs) encompass a broad spectrum of pathological conditions in which persistent or dysregulated inflammatory responses serve as major drivers of disease development and progression.1–3 Inflammation itself is an essential physiological process, acting as a frontline defense to eliminate infectious agents, repair tissue damage, and restore homeostasis following injury or infection.4,5 This process is typically characterized by a well-orchestrated sequence of events, beginning with rapid onset, a subsequent pro-inflammatory phase, and ultimately, a transition to resolution that supports tissue healing and functional recovery.6,7 However, disruptions in this finely tuned response, such as excessive activation or failure to resolve inflammation, can result in chronic or misdirected immune activity, which is central to the pathogenesis of a variety of IDs.8 Examples of such disorders include inflammatory bowel disease (IBD), hepatitis, atherosclerosis (AS), hypertension, multiple sclerosis (MS), uveitis, sensorineural hearing loss (SNHL), and neurodegenerative diseases.9 These conditions, though clinically diverse, display common pathological hallmarks: sustained immune activation, persistent oxidative stress, and an imbalance in tissue homeostasis that leads to ongoing injury and impaired function.10

At the molecular level, chronic inflammation is often triggered and perpetuated by the continuous presence of danger signals, either from invading pathogens (pathogen-associated molecular patterns (PAMPs)) or from host tissue damage-associated molecular patterns (DAMPs). These signals activate innate immune cells, prompting them to release a range of mediators including cytokines, chemokines, proteases, and reactive oxygen species (ROS), which together propagate the inflammatory cycle.11,12 Under normal circumstances, such immune responses are essential for clearing threats and restoring tissue integrity. However, when these mechanisms become excessive or dysregulated, they can initiate a self-amplifying cascade, resulting in widespread tissue damage, decreased immune tolerance, and ultimately the progression of inflammatory diseases.13,14 Understanding the dynamic relationship between inflammation and tissue homeostasis has paved the way for targeted strategies to modulate pathological inflammation. Conventional anti-inflammatory therapies such as corticosteroids, nonsteroidal anti-inflammatory drugs (NSAIDs), and biological agents targeting key cytokines remain the mainstay of treatment. Nevertheless, these therapeutic approaches are frequently limited by low specificity, rapid elimination, and potential off-target effects, often requiring high doses that may compromise safety and efficacy.15 These limitations highlight the need for more precise and innovative interventions, including emerging approaches that leverage nanobiotechnology and stimuli-responsive biomaterials, to achieve more effective management of IDs.

To address these limitations in current anti-inflammatory therapies, we propose a novel therapeutic paradigm termed pancatalysis, which fundamentally differs from traditional catalysis in both scope and mechanism.16 While conventional catalysis focuses primarily on accelerating isolated chemical reactions under controlled conditions, pancatalysis is realized through the development of advanced biomaterials that synergistically integrate three key components to achieve physiological modulation: (1) catalytic biomaterials are engineered to catalyze biochemical reactions triggered by endogenous stimuli (e.g., H2O2, glucose, glutathione, lactic acid) or external stimuli (e.g., light, ultrasound, electric/magnetic fields);17 (2) catalytic biology involves modulating biological molecules (e.g., proteins, lipids, nucleic acids) and pathways (e.g., oxidative stress, cell death, metabolism) using catalytic biomaterials;18 and (3) catalytic medicine focuses on precisely regulating reactions at disease sites for targeted therapy and diagnostics with minimal side effects, contrasting with traditional catalysis’ lack of therapeutic specificity.19,20 The operational foundation of this strategy is the P–A–N framework (preparation–activation–nontoxic treatment), which coordinates material design with biological response to achieve multi-level therapeutic outcomes, representing a level of systemic integration that conventional catalytic systems typically lack as they focus primarily on reaction efficiency.

Pancatalytic biomaterials, constituting both the operational foundation and physical embodiment of the pancatalysis paradigm, herald a transformative paradigm shift in ID treatment. Pancatalytic biomaterials, ranging from inorganic biomaterials to organic systems and even to organic/inorganic hybrid types, are engineered to trigger the specific biological effects by initiating catalytic chemical reactions, ultimately achieving nontoxic disease treatment. For instance, V2C MXenes exhibit multi-enzyme catalytic properties, neutralizing ROS and restoring redox balance in neurodegenerative models,21 while melanin-based composites resolve myocardial infarction (MI) microenvironments through proton-coupled electron transfer.22 By harnessing disease-specific stimuli such as pathological pH, hypoxia, or ROS overproduction, these catalytic biomaterials activate therapeutic cascades only at target sites, minimizing off-target effects. Central to this approach is the “Preparation of the catalyst, Activation of the biological effect and Nontoxic treatment of diseases” (P–A–N), which guides the design, fabrication and biomedical application of pancatalytic biomaterials.16,23 The P–A–N framework is exemplified by engineered GaHCF nanoabsorption catalysts (GaHCF NACs) for AS and CaSi2 nanosheets for rheumatoid arthritis (RA).24,25 In AS, GaHCF's preparation optimizes iron-binding affinity, while its activation phase sequesters pathologic iron and scavenges ROS to resolve ferroptosis–inflammation cycles; nontoxic treatment repurposes iron for diagnostic signaling. For RA, the preparation of calcium disilicide nanoparticles (CSNs) involves liquid exfoliation, while their activation phase hydrolyzes them in inflamed joints to release H2 for ROS scavenging and form bone-repairing calcium phosphate. The nontoxic treatment exhibits excellent biocompatibility and therapeutic performance in anti-inflammation and bone mineralization. These examples illustrate how the P–A–N framework enables the development of sophisticated therapeutic platforms capable of addressing the multifaceted nature of inflammatory diseases. By integrating material design with biological response modulation, pancatalytic biomaterials represent a significant advancement over conventional anti-inflammatory therapies, offering targeted, multi-modal intervention against the complex pathophysiology of inflammation.

This review systematically examines the emerging paradigm of pancatalytic biomaterials for inflammatory disease treatment through a multidisciplinary lens integrating materials science, biology and clinical medicine. We present a detailed design and classification of pancatalytic biomaterials based on their catalytic mechanisms and composition, with particular focus on their design principles under the P–A–N framework. The review thoroughly investigates their therapeutic applications across multiple physiological systems affected by chronic inflammation, such as cardiovascular, neurological, respiratory and sensory systems (Fig. 1). Special emphasis is placed on elucidating the structure–function relationships that enable these materials to simultaneously target oxidative stress, immune dysregulation and tissue damage, the key pathological drivers of inflammatory diseases. Finally, we critically analyze current translational challenges and outline future research directions to accelerate the clinical implementation of these innovative biomaterials for inflammatory disease management.


image file: d5nh00408j-f1.tif
Fig. 1 Schematic representation of pancatalytic biomaterials classified as inorganic, organic, and organic/inorganic hybrids, highlighting their therapeutic applications for inflammatory diseases across multiple organ systems, including cardiovascular, nervous, respiratory, digestive, joint, integumentary, sensory, urinary, and reproductive systems. Created with https://BioRender.com.

2. Design and classification of pancatalytic biomaterials

Pancatalytic biomaterials are a class of engineered materials designed to catalyze specific biochemical reactions, thereby modulating inflammatory pathways and providing targeted therapeutic interventions for IDs. These materials are classified into three main categories based on their composition and properties: inorganic, organic, and organic/inorganic hybrid catalytic biomaterials.

2.1 Design of pancatalytic biomaterials

Catalyzing specific biochemical reactions. The emerging catalytic biomaterials are designed to respond to both endogenous biological signals and external stimuli, thereby catalyzing specific biochemical reactions for personalized therapeutic interventions.17 Of these catalytic functions, scavenging ROS is particularly crucial in inflammatory diseases where oxidative stress plays a pathogenic role. Pancatalytic biomaterials are frequently engineered to mimic the catalytic activities of endogenous enzymes such as superoxide dismutase (SOD), catalase (CAT), and glutathione peroxidase (GPx), or are designed to directly scavenge ROS, which are critical for redox balance and inflammation control. By tailoring nanoscale structure and surface chemistry, these materials exhibit enzyme-like kinetics, enabling efficient scavenging of ROS in pathological microenvironments. Compared to natural enzymes, artificial nanozymes offer greater stability and sustained activity, providing targeted redox modulation and attenuating inflammation-associated oxidative stress, thus holding promise for advanced therapeutic interventions. For example, hexacyanoferrate nanocatalysts (CaH NCs) extend this paradigm by integrating multi-enzyme mimetic properties (SOD, CAT, GPx, and peroxidase) to regulate mitochondrial dysfunction and neuroinflammation in autism spectrum disorder.26 Similarly, amyloid fibril-templated ceria nanozymes demonstrate cascaded SOD-CAT-like catalysis with optimized Ce3+/Ce4+ ratios, synergistically alleviating oxidative stress, hypoxia, and hyperinflammation in diabetic wounds through glucose oxidase-coupled hydrogel microneedles.27
Biocompatibility and biodegradability. Ensuring the safety of biomaterials requires that these materials either undergo biodegradation into nontoxic byproducts or remain biologically inert within the body. Biocompatibility refers to the absence of cytotoxicity or adverse immune responses, such as inflammation or allergy, from both the intact material and its degradation products. Biodegradability allows materials to be broken down under physiological conditions into substances that can be efficiently cleared, thereby reducing the risk of bioaccumulation and minimizing long-term toxicity. For materials that are non-degradable, chemical stability and minimal biological reactivity are essential to avoid undesired interactions with tissues and organs. These criteria form the basis for evaluating the clinical applicability of novel nanomaterials. For instance, pH-responsive MoO3−x nanourchins rapidly biodegrade into renal-clearable molybdate ions under physiological conditions (pH 7.4), eliminating off-target catalytic activity in healthy tissues while retaining tumor-specific enzymatic effects in acidic microenvironments.28 Meanwhile, zinc sulfide-based nanozymes leverage the intrinsic biocompatibility of zinc, degrading to release thiolate ligands that accelerate bioorthogonal catalysis while ensuring metabolic clearance of zinc ions.29 These designs exemplify how controlled degradation pathways mitigate bioaccumulation risks and align with the imperative for clinically translatable, safety-first nanotherapies.
Targeting and activation mechanisms. The precision of anti-inflammatory therapy is typically enhanced by employing pancatalytic biomaterials engineered for targeted delivery and stimuli-responsive activation. Such materials are designed to accumulate preferentially at pathological sites based on physicochemical properties, including particle size, surface charge, and the presence of affinity ligands.30 Therapeutic activity is then selectively triggered by specific microenvironmental cues, such as acidic pH, elevated ROS, or increased expression of certain enzymes. This strategy improves local efficacy while reducing systemic side effects, contributing to the optimization of therapeutic outcomes in inflammatory diseases.31,32 For example, in IBD therapy, inflamed colon-targeted antioxidant nanotherapeutics are constructed by functionalizing mesoporous silica nanoparticles (MSNs) with ROS-scavenging ceria nanoparticles (Ce NPs), followed by poly(acrylic acid) coating to enable electrostatic targeting of inflamed colon tissues.33 In MI, Fe-Cur@TA nanozymes are modified with tannic acid, enabling them to specifically bind to cardiac elastin and collagen for preferential localization in damaged heart regions.34 These targeting strategies enhance local drug concentration and allow microenvironment-responsive activation, thereby improving therapeutic precision in various diseases.
Multifunctionality. Pancatalytic biomaterials bring together various functions, combining therapy, diagnosis, and regeneration into one cohesive platform. This holistic approach enables better disease management, where real-time monitoring of therapeutic actions is possible, and regenerative processes can be supported effectively. By utilizing these combined properties, these biomaterials can target specific pathological signals and help track the effectiveness of treatments, thereby improving clinical results. This multifunctionality marks an important step forward in biomedical engineering, offering new ways to tackle the complexities of managing diseases compared to traditional therapies. For instance, in treating acute liver failure, researchers have developed a ROS-sensitive nanozyme-augmented photoacoustic nanoprobe that allows for early diagnosis and timely nanocatalytic therapy.35 This innovative system pairs the imaging capabilities of near-infrared agents with the catalytic actions of ceria nanozymes, enabling real-time monitoring of ROS and addressing oxidative stress at the injury site simultaneously. In a similar vein, using Mn3O4 nanozymes to engineer stem cells for MI therapy boosts their therapeutic impact by enhancing ROS scavenging capabilities and allows for tracking via magnetic resonance imaging.36 These multifunctional strategies illustrate how pancatalytic biomaterials can synergize various treatment approaches, leading to more effective healthcare solutions across different diseases.

2.2 Classification of pancatalytic biomaterials

In the field of anti-inflammatory therapy, pancatalytic biomaterials have attracted significant attention due to their ability to precisely modulate disease microenvironments, catalytically eliminate ROS, and regulate immune responses. Based on differences in composition and functional properties, these materials are mainly classified into three categories: inorganic, organic, and organic/inorganic hybrid systems. Inorganic catalytic materials, with their outstanding multi-enzyme-mimetic activities and stability, exhibit remarkable advantages in antioxidation and inflammation regulation; organic materials emphasize biocompatibility and structural tunability, often achieving efficient anti-inflammatory effects by mimicking natural enzyme functions or through self-assembly; and hybrid systems integrate the catalytic activity of inorganic components with the adaptability of organic constituents, enabling multifunctional and precise regulation. The following sections will discuss representative materials from each category of pancatalytic biomaterials and recent advances in their applications for IDs.

Inorganic catalytic biomaterials are widely studied due to their exceptional catalytic efficiency, stability, and tunable redox properties. Cerium oxide (CeO2) nanoparticles exemplify this category, mimicking SOD and CAT activities through Ce3+/Ce4+ redox cycling. These nanoparticles effectively scavenge ROS in IBD and MI models.22,33 Prussian blue analogs (e.g., GaHCF) further demonstrate therapeutic potential by sequestering excess iron in atherosclerotic plaques, disrupting ferroptosis–inflammation interactions while enabling non-invasive photoacoustic imaging.24 Advanced materials like V2C MXenes exhibit multi-enzyme mimetic properties (SOD, CAT, GPx), providing broad-spectrum antioxidant protection in neurodegenerative and inflammatory diseases.21 Additionally, black phosphorus nanosheets (BPNSs) catalyze ROS degradation to promote tissue repair.37

Organic catalytic biomaterials comprise both purely organic systems and organic-based hybrid constructs, such as natural polyphenols (including quercetin, resveratrol, and tannic acid), synthetic conducting polymers (like polypyrrole and polyaniline), polysaccharide derivatives (e.g., chitosan, alginate), artificial antioxidant polymers, supramolecular organic frameworks, and engineered enzyme systems like antioxidant enzyme-loaded polymersomes. These materials are engineered to mimic natural enzyme activities or exhibit direct ROS-scavenging capabilities, regulate redox homeostasis, and facilitate targeted drug delivery, making them highly versatile for biomedical applications.38 These materials bridge the gap between natural antioxidant enzymes (SOD, CAT) and chemical antioxidants (vitamin C), combining their advantages while overcoming limitations such as poor enzyme stability and transient antioxidant effects. Moreover, engineered modifications of natural enzymes can effectively compensate for their inherent limitations. Recent advances include ROS-responsive polyion complex vesicles (PICsomes) encapsulating catalase, where engineered semipermeable membranes enhance enzyme stability and enable sustained redox regulation through cysteamine-mediated GSH biosynthesis. The results further demonstrate significant long blood circulation of PICsomes loaded with catalase and strong protection effects against bloodstream oxidative stress, paving the way for the further development of truly effective in vivo therapeutics.39 This innovative approach not only addresses the stability issues of natural enzymes but also provides sustained catalytic activity and targeted delivery, representing a significant advancement in biomimetic antioxidant therapy. A notable example is the development of CAT-loaded nanogels (CAT-NGs) for neutrophilic asthma treatment, where chitosan-based nanogels significantly improved CAT stability and bioavailability while endowing additional antibacterial properties. This system demonstrated enhanced therapeutic efficacy in alleviating airway inflammation and oxidative stress through efficient ROS scavenging and inhibition of NOD-like receptor family pyrin domain containing 3/nuclear factor kappa-light-chain-enhancer of activated B cells (NLRP3/NF-κB) pathways.40 For chemical antioxidants like vitamin C, vitamin C–lipid nanoassemblies effectively address IBD treatment challenges. Through fatty acid conjugation, these self-assembling nanoparticles enhance vitamin C's stability and enable targeted delivery to inflamed colon tissues. In mouse models, they demonstrate significant ROS scavenging and disease mitigation. When combined with budesonide, they provide synergistic anti-inflammatory effects while maintaining immune function and gut microbiome balance, overcoming vitamin C's inherent pharmacokinetic limitations.41 These examples collectively demonstrate how advanced material engineering can optimize both enzymatic and non-enzymatic antioxidant systems for improved therapeutic outcomes.

Among them, polydopamine (PDA) and epigallocatechin gallate (EGCG) are particularly prominent. PDA, inspired by mussel adhesive proteins, possesses abundant catechol groups that endow it with CAT- and SOD-like activities, enabling effective scavenging of ROS, modulation of inflammatory responses, and promotion of tissue repair. PDA-based platforms have demonstrated significant protective effects in neural injury and degenerative disease models by reducing oxidative stress and enhancing neuronal survival.42,43 EGCG, a key polyphenol found in green tea, can self-assemble or co-assemble into nanoparticles, substantially improving its inherent antioxidant and anti-inflammatory capabilities. When incorporated into advanced delivery systems, EGCG-based nanomaterials have shown efficacy in alleviating radiation-induced skin injury, suppressing tissue inflammation, and accelerating wound healing.44 Together, these examples demonstrate the broad therapeutic promise of organic catalytic biomaterials in precisely regulating oxidative stress and inflammation, and in promoting the repair of damaged tissues across various disease models. Their structural diversity and multifunctionality highlight their potential for future clinical translation in the treatment of inflammatory, neurodegenerative, and other oxidative stress-related conditions.

Hybrid systems are engineered to combine the robust catalytic activity and redox properties of inorganic materials with the structural versatility, biocompatibility, and functional tunability of organic components, enabling multifunctional and highly precise therapeutic interventions. By integrating these complementary characteristics, such systems can achieve enhanced catalytic efficiency, controlled drug delivery, responsive targeting, and adaptive interactions with complex biological environments.38 MOFs, such as Ce-MOFs, deliver dexamethasone while scavenging ROS in uveitis, leveraging their porous structures for controlled drug release.45 Biomimetic nanosystems like Fe-Cur@TA nanozymes merge Fe3+–curcumin coordination for ROS scavenging with tannic acid-mediated cardiac targeting, effectively disrupting the inflammation-free radical cycle in MI.34 Similarly, melanin-based composite nanomedicines (MCN) integrate PDA, Prussian blue, and CexOy to neutralize ROS cascades in ischemic heart disease, showcasing synergistic catalytic and regenerative effects.22

3. Applications of pancatalytic biomaterials in inflammation-related disease

The remarkable versatility of pancatalytic biomaterials is demonstrated by their broad therapeutic applications across multiple physiological systems affected by inflammatory diseases, as systematically summarized in Table 1. These advanced materials, categorized as inorganic, organic, and hybrid systems, have been successfully engineered to target specific inflammatory pathologies through their unique catalytic properties. The table highlights representative examples of pancatalytic biomaterials that have shown efficacy in diverse disease models, ranging from cardiovascular disorders (e.g., GaHCF NACs for atherosclerosis) to neurological conditions (e.g., IP6@GCA/AP for Alzheimer's disease), respiratory diseases (e.g., CAT-NGs for neutrophilic asthma), and beyond. In the following sections, each representative example listed in the table will be analyzed through the lens of the P–A–N framework, with detailed discussions on how the preparation, activation, and nontoxic treatment components are specifically engineered for different disease applications.
Table 1 Representative pancatalytic biomaterials for inflammatory diseases across physiological systems, including material classification, animal models, and targeted diseases
System Pancatalytic biomaterials Classification Animal model Type of IDs Ref.
Cardiovascular GaHCF NACs Inorganic High-fat diet-fed ApoE/− mice Atherosclerosis 24
Cardiovascular MCN Organic/inorganic MI model in mice Myocardial infarction 22
Cardiovascular Pt@LF Organic/inorganic Ischemic stroke in mice Stroke 46
Cardiovascular CaHF NPs Inorganic Stress-induced hypertension in mice Hypertension 47
Nervous IP6@GCA/AP Organic/inorganic 5 × FAD mice Alzheimer's disease 48
Nervous ATB NPs Organic/inorganic Preformed fibril induced PD mice Parkinson's disease 49
Nervous MnFe2O4 NPs Inorganic HD model in mice Huntington's disease 50
Nervous MSN-CeNPs Inorganic Experimental autoimmune encephalomyelitis in mice Multiple sclerosis 51
Respiratory PEG@CS/BPQDs-AM NPs Organic/inorganic The mice after cigarette smoking and P. aeruginosa infection Chronic obstructive pulmonary disease 52
Respiratory CAT-NGs Organic Mice infected with NTHi and subjected to OVA Neutrophilic asthma 40
Digestive RuCo nanosheets Inorganic DSS-induced IBD model in mice Inflammatory bowel disease 53
Digestive mRNA LNPs Organic Liver fibrosis mouse models Metabolic dysfunction-associated steatohepatitis 54
Joint CaSi2 Inorganic Adjuvant-induced arthritis mice Rheumatoid arthritis 25
Joint BPNSs Inorganic ACLT in rats Osteoarthritis 37
Integumentary Cu/Zn-MOF Organic/inorganic MC903-induced AD-like mouse model Atopic dermatitis 55
Integumentary FeN4O2-SACs Inorganic IMQ-induced psoriasiform model in mice Psoriasis 56
Sensory Gel/Mn/NAC/PDA Organic/inorganic Cisplatin-induced ototoxicity model in mice Sensorineural hearing loss 57
Sensory Cs@P/CeO2 Organic/inorganic BAK-induced DED model in mice Dry eye disease 45
Sensory DSP@Ce-MOFs F127 hydrogel Organic/inorganic Endotoxin-induced uveitis in mice Uveitis 58
Sensory MLPGa Organic/inorganic C. albicans-infected fungal keratitis model in mice Fungal keratitis 59
Urinary Pt5.65S Inorganic Cisplatin-induced AKI in mice Acute kidney injury 60
Urinary DEC Organic/inorganic CFT073-induced UTIs in mice Urinary tract infections 61
Reproductive FeLab Organic/inorganic C. albicans-induced vaginitis in mice Candida vaginitis 62
Reproductive TPDA NPs Organic LPS-induced acute prostatitis mouse models Prostatitis 63
Others DNA-pDHN Organic 4T1 tumor-bearing mouse models Tumor 64
Others UPDA NPs Organic DOX-induced senescence in mice Aging 65


3.1 Cardiovascular system

Cardiovascular diseases (CVDs), encompassing atherosclerosis, thrombosis, myocardial/cerebral infarction, and subsequent ischemia-reperfusion injury, represent a major global health threat.66–68 According to the World Health Organization (WHO), CVDs account for approximately 17.9 million deaths annually worldwide.69 These conditions not only drive high mortality but also cause severe disabilities, such as heart failure following MI or motor dysfunction post-stroke, imposing a profound socioeconomic burden. CVDs encompass ischemic or hemorrhagic disorders affecting the brain, heart, and systemic vasculature, predominantly arising from multifactorial etiologies such as AS, hypertension, dyslipidemia, and increased blood viscosity.67 Clinically, these conditions manifest with elevated rates of morbidity, mortality, disability, and recurrence, frequently resulting in multiorgan complications.

AS, characterized by chronic inflammation and vascular dysfunction, underlies the pathogenesis of CVDs, and is marked by the formation of lipid-rich plaques on arterial walls.69,70 These plaques arise from lipid deposition and inflammatory activation, initiating with endothelial dysfunction that permits low-density lipoprotein (LDL) infiltration into the subendothelial space. Oxidative modification converts LDL to oxidized LDL (ox-LDL), triggering monocyte recruitment and differentiation into macrophages.71 Macrophages engulf ox-LDL to form foam cells, which accumulate in the arterial intima, perpetuating chronic inflammation and releasing cytokines such as interleukin-6 (IL-6) and tumor necrosis factor-α (TNF-α).72,73 Subsequently, vascular smooth muscle cells migrate to the intimal layer, undergo aberrant proliferation, and secrete extracellular matrix components such as collagen and elastin, collectively contributing to the formation and stabilization of fibrous caps. As lipid cores expand, inflammatory cell infiltration intensifies, and fibrous tissue hyperplasia progresses, arterial walls undergo progressive thickening and calcification, culminating in luminal narrowing and the development of mature atherosclerotic plaques.74 Notably, rupture or erosion of vulnerable plaques may trigger thrombus formation, leading to acute vascular occlusion and subsequent ischemic events such as MI or stroke, which remain the leading causes of global mortality. Current clinical strategies primarily focus on restoring blood flow via thrombolysis or percutaneous coronary intervention. However, rapid reperfusion often induces oxidative stress-mediated tissue damage, a phenomenon termed ischemia-reperfusion injury, which paradoxically complicates therapeutic outcomes.75–77 Hypertension induces endothelial damage through mechanical stress (such as elevated shear forces) and biochemical pathways (such as overactivation of the renin–angiotensin system). This exacerbates the progression of CVD and thus accelerates atherosclerosis.78

In response to these challenges, new biomaterials, particularly pancatalytic biomaterials, show innovative potential. By integrating multi-modal response characteristics (such as pH, ROS, angiotensin converting enzyme), these materials can accurately identify the pathological microenvironment of CVD and achieve integrated diagnosis and treatment.

AS, a chronic inflammatory disease characterized by lipid-rich plaque formation in arterial walls, poses a significant global health burden due to its progression to life-threatening cardiovascular events such as MI and stroke.79–81 Central to its pathology is the development of unstable “vulnerable plaques”, which feature thin fibrous caps, necrotic cores enriched with oxidized lipids, and intraplaque hemorrhage. These plaques are prone to rupture, driven by the cell death mechanism and sustained inflammation.82,83 Despite advances in pharmacotherapy, current strategies including iron chelators (e.g., deferoxamine), ferroptosis inhibitors (e.g., ferrostatin-1), and anti-inflammatory biologics (e.g., canakinumab) face critical limitations.84–87 High costs, systemic toxicity (e.g., increased infection risk with interleukin inhibitors), and the need for polypharmacy to compensate for single-target inefficacy hinder clinical translation. Moreover, conventional therapies fail to address the spatial coexistence of cell death, oxidative stress, and chronic inflammation within plaques, underscoring the need for multifunctional interventions.88 Pancatalytic biomaterials address these challenges through integrated preparation (P), activation (A), and nontoxic treatment (N). Engineered GaHCF NACs exemplify this approach, leveraging their catalytic properties to achieve multifaceted therapeutic effects. During preparation, GaHCF NACs are designed with gallium's iron-mimetic properties to ensure structural stability and high iron-binding affinity. In the activation phase, they dynamically capture excess iron and scavenge reactive oxygen/nitrogen species (RONS), disrupting ferroptosis and lipid peroxidation while repolarizing macrophages toward anti-inflammatory M2 phenotypes. This dual catalytic action reduces necrotic core expansion and enhances fibrous cap stability. The nontoxic treatment phase leverages GaHCF NACs’ in situ photoacoustic signaling capability, which converts sequestered iron into diagnostic signals for real-time plaque monitoring. By neutralizing pathologic iron while utilizing it for theranostics, GaHCF NACs circumvent systemic toxicity and enhance biocompatibility (Fig. 2(a)). This tripartite strategy of synergistic material design, catalytic microenvironment modulation, and self-monitoring functionality demonstrates how pancatalysis holistically resolves the iron–inflammation–oxidation axis, offering a paradigm for precision treatment of inflammatory CVDs.24


image file: d5nh00408j-f2.tif
Fig. 2 Pancatalytic biomaterials for the treatment of cardiovascular disease. (a) GaHCF NACs for AS treatment via the P–A–N framework. Reproduced with permission.24 Copyright 2025 American Chemical Society. (b) Diagram illustrating the synthesis process of MCN. Reproduced with permission.22 Copyright 2024, Springer Nature. (c) Diagrammatic representation emphasizing the role of MCN in ameliorating hypoxic conditions. Reproduced with permission.22 Copyright 2024, Springer Nature. (d) Representative images of HE staining on day 28 after MI. Reproduced with permission.22 Copyright 2024, Springer Nature. (e) Pt@LF nanomotor for stroke treatment via the P–A–N framework. Reproduced with permission.46 2024 Wiley-VCH GmbH. (f) CaHF NPs for hypertension treatment via the P–A–N framework. Reproduced with permission.47 Copyright 2023 Elsevier Ltd. All rights reserved.

MI, a critical manifestation of CVDs, remains a leading cause of global mortality. Despite advancements in preventive strategies, approximately 50 million individuals suffer from MI annually.68,76,80 The pathological progression of MI is tightly regulated by the dynamic myocardial infarction microenvironment (MIM), a tripartite system dominated by hypoxia, acidosis (pH 6.5–6.8), and redox imbalance due to excessive ROS.89–91 These interconnected perturbations disrupt cardiomyocyte survival pathways, exacerbate mitochondrial dysfunction, and drive inflammatory cascades, ultimately leading to fibrotic scarring and irreversible cardiac damage.92–94 Current clinical interventions, such as reperfusion therapy, paradoxically aggravate oxidative stress and fail to prevent ischemia-reperfusion injury, underscoring the urgent need for novel therapeutic paradigms.95 Pancatalytic biomaterials break this deadlock through holistic catalytic management spanning preparation (P), activation (A), and nontoxic treatment (N). MCN exemplifies this enzyme-inspired catalytic paradigm. During preparation, MCN synergistically assembles PDA, Prussian blue, and cerium oxide into a multi-catalytic nanoarchitecture mimicking enzymatic cooperativity (Fig. 2(b)). In the activation phase, it employs proton-coupled electron transfer and redox catalytic cycling to convert pathologic H+, ROS, and hypoxia into O2 and H2O through self-sustained catalytic cascades, simultaneously resolving acidosis, quenching oxidative stress, and restoring oxygen homeostasis (Fig. 2(c)). These catalytic actions dismantle the MIM's vicious cycle while polarizing macrophages toward anti-inflammatory phenotypes. The nontoxic treatment phase leverages MCN's endogenous components and clinically validated materials to ensure localized metabolically catalytic activity and safe clearance, eliminating off-target risks (Fig. 2(d)).22

Stroke ranks as the second leading global cause of death and the third contributor to chronic disability, with half of survivors facing lifelong functional impairment.96 Projections indicate a 1.5-fold increase in stroke-related mortality by 2050 compared to 2020, driven by aging populations and insufficient therapeutic advancements.79,97,98 The ischemic penumbra, a salvageable region surrounding the irreversibly damaged core, faces dual threats: energy failure from hypoperfusion and “no-reflow” phenomena caused by microthrombi and pericyte-mediated vasoconstriction, which impede drug delivery. Current clinical strategies rely on thrombolytics like recombinant tissue-type plasminogen activator (rt-PA), but their narrow therapeutic window (<4.5 hours) limits applicability, while delayed interventions (e.g., antithrombotics or neuroprotectants) exhibit diminished efficacy.99,100 Clinical limitations stem from three interrelated barriers: (1) inadequate penetration of therapeutics into hypoxic regions due to microvascular occlusion and blood–brain barrier (BBB) dysfunction;101 (2) the inability of monotherapies (e.g., ROS scavengers or iron chelators) to address the ferroptosis–excitotoxicity axis;102 and (3) systemic toxicity risks from non-targeted drug distribution.103 Pancatalytic biomaterials address these gaps through the P–A–N framework. The Pt@LF nanomotor exemplifies this paradigm (Fig. 2(e)). During preparation, platinum nanoclusters (Pt NCs) are functionalized with apo-lactoferrin (Apo-LF), combining Pt's dual CAT-like and NH4+ scavenging activities with Apo-LF's iron-chelating capacity and BBB-targeting capability via lactoferrin receptor binding. In the activation phase, Pt NCs convert pathologic ROS into O2, mitigating lipid peroxidation and restoring redox balance, while Apo-LF sequesters excess iron to inhibit ferroptosis drivers. Simultaneously, NH4+ clearance prevents N-methyl-D-aspartic acid receptors’ overactivation, breaking the excitotoxicity vicious cycle. The nanomotor's self-propelling motion enhances penetration through occluded microvessels, enabling mechanical thrombolysis and targeted delivery to the penumbra. For nontoxic treatment, Apo-LF's endogenous origin and Pt's metabolic stability ensure biocompatibility, minimizing off-target effects. By synchronously modulating iron homeostasis, redox imbalance, and neuroinflammation, Pt@LF reprograms the ischemic microenvironment, polarizing microglia toward anti-inflammatory phenotypes and reducing infarct volume.46

Hypertension, affecting over 1.28 billion individuals globally in 2019 and projected to rise to 1.56 billion by 2025, represents a critical public health challenge due to its strong association with cardiovascular, cerebrovascular, and renal complications.104,105 The pathogenesis is intricately linked to oxidative stress and neuroinflammation driven by RONS, including superoxide anion (˙O2), hydrogen peroxide (H2O2), hydroxyl radicals (˙OH), nitric oxide (NO), and peroxynitrite (ONOO).106–108 In the rostral ventrolateral medulla, a key regulator of sympathetic nervous activity, RONS overproduction originates from nicotinamide adenine dinucleotide phosphate (NADPH) oxidase (NOX) and mitochondrial dysfunction. These reactive molecules disrupt redox balance, amplify sympathetic excitability, and reduce NO bioavailability, converting its neuroprotective role into a cytotoxic one through ONOO formation under hypoxic conditions.109 Concurrently, microglia, the brain's resident immune cells, shift to a proinflammatory state in response to RONS, releasing cytokines that further exacerbate NOX-mediated ROS generation, DNA damage, and neuronal apoptosis. This self-reinforcing vicious cycle underscores the interplay between oxidative stress, neuroinflammation, and sympathetic overactivation in sustaining hypertensive pathology.110,111

Despite advances in understanding hypertensive pathophysiology, clinical management of hypertension continues to face significant challenges, including poor patient adherence to lifelong medication, drug-related side effects (e.g., hypotension, renal impairment), and the complexity of managing comorbidities (e.g., diabetes, chronic kidney disease). Additionally, limited efficacy of lifestyle interventions and the prevalence of resistant hypertension further complicate sustained blood pressure control.112 Pancatalytic biomaterials bridge these gaps through integrated preparation (P), activation (A), and nontoxic treatment (N). Calcium hexacyanoferrate (III) nanoparticles (CaHF NPs) exemplify this holistic catalytic system (Fig. 2(f)). During preparation, ultrasmall CaHF NPs are engineered to mimic SOD, CAT, GPx, and POD activities, creating a multi-enzyme catalytic capacity for broad-spectrum RONS neutralization. In the preparation phase, ultrasmall CaHF NPs are rationally engineered with multi-enzyme mimetic activities, including SOD, POD, GPx, and CAT, to enable efficient scavenging of diverse RONS. During the activation phase, these nanoparticles are triggered by the pathological microenvironment, where elevated RONS and neuroinflammation drive hypertension progression; CaHF NPs catalytically convert harmful species such as H2O2, ˙O2, ˙OH, and NO into harmless molecules, thereby alleviating oxidative stress, suppressing inflammation, and reducing neuronal apoptosis. In the nontoxic treatment phase, CaHF NPs demonstrate high biocompatibility and minimal side effects, overcoming the stability and safety limitations of natural enzymes and traditional nanozymes. CaHF NPs exemplify how concerted catalytic engineering resolves the interconnected drivers of hypertension.47

3.2 Nervous system

Neuroinflammation is a complex immune response of the central nervous system (CNS) to harmful stimuli such as infection, ischemia, or trauma.113 Its core characteristics include excessive activation of glial cells (microglia and astrocytes), elevated release of cytokines (pro-inflammatory cytokines, chemokines, and ROS), migration of peripheral immune cells, and localized tissue damage.114 Under physiological conditions, neuroinflammation maintains brain homeostasis by clearing pathogens and repairing damaged tissues. However, chronic or dysregulated inflammatory responses trigger neuronal injury and degeneration, serving as a central driving factor for neurodegenerative diseases such as Alzheimer's disease (AD), Parkinson's disease (PD), Huntington's disease (HD), and multiple sclerosis (MS).114

Neuroinflammation is a complex process driven by the interplay of central and peripheral immune components within the CNS. At its core, this mechanism involves the activation of resident glial cells, notably microglia and astrocytes, which dynamically regulate inflammatory responses. Microglia, the primary immune sentinels of the CNS, exhibit functional plasticity by polarizing into pro-inflammatory M1 phenotypes or anti-inflammatory M2 phenotypes. M1 microglia exacerbate neurotoxicity through excessive release of pro-inflammatory cytokines such as interleukin-1β (IL-1β) and TNF-α, along with ROS, while M2 microglia promote tissue repair via anti-inflammatory mediators like interleukin-10 (IL-10). Similarly, astrocytes adopt reactive states classified as neurotoxic A1 or neuroprotective A2 subtypes, influencing neuroinflammation through cytokine production, metabolic regulation, and BBB modulation. The infiltration of peripheral immune cells, including neutrophils, monocytes, and lymphocytes, further amplifies CNS inflammation. Neutrophils disrupt BBB integrity through the release of proteases and the formation of neutrophil extracellular traps, while monocytes differentiate into macrophages that interact with microglia to shape inflammatory outcomes. Cytokines such as IL-6, interleukin-23, and granulocyte-macrophage colony-stimulating factor act as critical signaling molecules, activating pathways like Janus kinase-signal transducer and activator of transcription (JAK-STAT) and mitogen-activated protein kinase (MAPK) to coordinate immune responses. Chronic dysregulation of these processes leads to sustained inflammation, contributing to neurodegenerative pathologies.

Neuroinflammation is characterized by a dynamic interplay of central and peripheral immune mechanisms triggered by pathogens, injury, or stress.113 Central to this process are microglia, the CNS-resident immune cells, which exhibit dual functional polarization into pro-inflammatory M1 or anti-inflammatory M2 phenotypes.115 M1 microglia, activated via Toll-like receptors (TLRs) and interferon-γ signaling, release cytotoxic mediators such as IL-1β, TNF-α, and ROS, exacerbating neuronal damage.116 Conversely, M2 microglia promote tissue repair through anti-inflammatory cytokines including IL-10 and transforming growth factor-beta (TGF-β), as well as various growth factors. Astrocytes similarly adopt reactive states, namely, neurotoxic A1 or protective A2 subtypes, regulated by pathways including NF-κB, MAPK, and JAK/STAT3.117,118 Reactive astrocytes influence neuroinflammation by modulating cytokine release, BBB integrity, and metabolic support, though chronic activation disrupts neuronal homeostasis and amplifies neurodegeneration.119–122 Peripheral immune cells, such as neutrophils and monocytes, infiltrate the CNS under pathological conditions.123 Neutrophils compromise the BBB via proteases and extracellular traps, while monocytes differentiate into macrophages that interact with microglia, shaping inflammatory outcomes.124,125 Cytokines like IL-6, IL-1β, and TNF, produced by glial cells and infiltrating leukocytes, activate signaling pathways (e.g., JAK-STAT, MAPK) to propagate inflammation.126 Persistent dysregulation of these interactions, marked by sustained cytokine release, glial hyperactivation, and immune cell influx, drives chronic neuroinflammation, contributing to neurodegenerative disorders.

The intricate pathological networks of neurodegenerative diseases, such as glial cell phenotypic imbalance and oxidative-inflammatory cycles, urgently require multi-target synergistic intervention strategies. Pancatalytic biomaterials, which integrate biocatalysis (e.g., enzyme-mimetic activity), material catalysis (e.g., ROS-scavenging nanozymes), and medical regulation (e.g., targeted delivery), achieve holistic management through the P–A–N framework. For example, AD as a representative neurodegenerative disorder is pathologically characterized by β-amyloid (Aβ) plaque deposition, neurofibrillary tangles caused by tau hyperphosphorylation, chronic inflammation, and neuronal apoptosis.127 While conventional single-target therapies like Aβ-targeting antibodies (e.g., Aducanumab) partially clear plaques, they fail to address the complex interplay of multiple pathological pathways in AD, underscoring the necessity for multi-target synergistic therapies.128 Recently, a supramolecular assembly-based platform strategy (IP6@GCA/AP) has achieved systemic intervention in AD pathology through dynamic modular design (Fig. 3(a)). Key mechanisms include the following: guanidinium-modified calixarene (GCA) inhibits fibril formation and disassembles pre-existing fibrils via strong interactions between guanidinium groups and negatively charged residues (aspartate/glutamate) on Aβ42; its low hydration energy enhances microglial phagocytosis and clearance of Aβ. Ascorbyl palmitate (AP), acting as an antioxidant, alleviates oxidative stress in the AD brain by scavenging ROS, thereby mitigating inflammation; its hydroxyl groups form heteromultivalent bonds with GCA, enhancing binding affinity to Aβ. Dipotassium phytate (IP6), precisely released via a biomarker displacement activation strategy, inhibits β-secretase activity to reduce Aβ production.48


image file: d5nh00408j-f3.tif
Fig. 3 Pancatalytic biomaterials for the treatment of nervous system diseases. (a) IP6@GCA/AP for AD treatment via the P–A–N framework. Reproduced with permission.48 Copyright 2024 Wiley-VCH GmbH. (b) ATB NPs for PD treatment via the P–A–N framework. Reproduced with permission.49 Copyright 2025, The American Association for the Advancement of Science. (c) MnFe2O4 NPs for HD treatment via the P–A–N framework. Reproduced with permission.50 Copyright 2019 Elsevier Ltd. All rights reserved. (d) Energy dispersive X-ray spectroscopy elemental mapping. Reproduced with permission.51 Copyright 2022, Springer Nature. (e) Scheme demonstrates the catalytic property of CeNPs to scavenge intracellular ROS for the suppression of APC activation. Reproduced with permission.51 Copyright 2022, Springer Nature. (f) The number and frequency of CD3+CD4+ T cells infiltrating the spinal cord and the frequency of CD3+CD4+ T cells in cervical lymph nodes on day 31. Reproduced with permission.51 Copyright 2022, Springer Nature.

PD is a neurodegenerative disorder centered on the degeneration of dopaminergic neurons in the substantia nigra pars compacta and striatal dopamine (DA) depletion.129 Clinically manifested as motor dysfunction and cognitive impairment, its pathology is closely associated with α-synuclein (α-syn) aggregates forming Lewy bodies, which directly induce neuronal toxicity and exacerbate pathological progression by suppressing autophagy systems such as chaperone-mediated autophagy (CMA).129–131 Current PD treatments, such as levodopa and DA agonists, alleviate symptoms by boosting DA levels but cannot regenerate degenerated neurons or enable precise spatial modulation in the substantia nigra, failing to halt disease progression.132,133 While deep brain stimulation (DBS) activates neurons via external stimuli (e.g., near-infrared light), it necessitates permanent device implantation, risking infections and hardware complications.134 Optogenetic DBS alternatives avoid implants but rely on viral-mediated gene delivery for light-sensitive proteins, raising safety issues linked to genetic manipulation.135 To address this, researchers developed a wireless photothermal-driven DBS nanosystem (Au@TRPV1@β-syn NPs, ATB NPs). ATB NPs exemplify the P–A–N paradigm by synergizing photothermal energy conversion, targeted biological modulation, and protein homeostasis restoration (Fig. 3(b)). The nanosystem is composed of gold nanoshells (AuNSs) for near-infrared (NIR, 808 nm) light-to-heat conversion, transient receptor potential vanilloid family member 1 (TRPV1) antibodies conjugated to AuNSs to anchor specifically to DA neurons in the substantia nigra (SN), and β-synuclein (β-syn) peptides linked via NIR-responsive bonds for competitive binding to the hydrophobic regions of α-syn. Following stereotaxic delivery to the SN, ATB NPs localize on DA neurons through TRPV1 targeting. Upon pulsed NIR irradiation, the nanosystem acts as a nanoantenna to convert light into localized heat, which simultaneously activates TRPV1 channels to induce Ca2+ influx and restore DA neuronal activity while triggering the release of β-syn peptides to disrupt α-syn fibrils. The heat further stimulates CMA to enhance lysosomal clearance of pathological α-syn preformed fibrils. This spatiotemporally controlled process elevates striatal DA levels, degrades toxic aggregates, and reverses motor deficits in PD mice, achieving these outcomes without the invasive risks associated with DBS. By integrating physical energy transduction with biological pathway regulation, ATB NPs demonstrate the transformative potential of pancatalytic biomaterials in precision neuromodulation and neurodegenerative disease therapy.49

HD is an autosomal dominant neurodegenerative disorder caused by abnormal polyglutamine (polyQ) expansion (CAG repeats >35) in the N-terminal region of huntingtin (Htt) protein.136 Its core pathology involves toxic aggregation of mutant Htt (mHtt), leading to neuronal dysfunction, motor impairments, short-term memory deficits, and progressive death.137 MnFe2O4 nanoparticles (NPs) exemplify a pancatalytic biomaterial strategy for HD therapy through the P–A–N framework. During the preparation phase, biocompatible MnFe2O4 NPs are synthesized via a simple method and delivered to neuronal cells stably expressing mutant huntingtin with 74 polyglutamine repeats (mHtt-Q74). In the activation phase, these NPs selectively enhance K48-linked ubiquitination of mHtt-Q74, directing its recognition by the ubiquitin receptor ubiquilin-1 (not p62/SQSTM1) for targeted degradation via the ubiquitin-proteasome system (UPS), bypassing autophagy-lysosomal pathways. This UPS-specific activation overcomes prior limitations of nanomaterials relying solely on autophagy. In the nontoxic treatment phase, MnFe2O4 NPs achieve precise clearance of toxic mHtt-Q74 aggregates without compromising wild-type protein function or inducing cytotoxicity, effectively restoring proteostasis (Fig. 3(c)). This work pioneers the use of nanomaterials to exploit UPS-mediated degradation for HD, offering a novel therapeutic avenue for polyglutamine expansion disorders.50

MS is a neuroinflammatory autoimmune disorder characterized by autoreactive CD4+ T cell-mediated demyelination and axonal damage in the CNS, leading to progressive motor dysfunction.138 Existing treatment regimens for MS primarily achieve transient alleviation of chronic neuropathic symptoms while carrying the risk of systemic immune compromise.139 To address this, a pancatalytic biomaterial-based immunosuppressive vaccine (MSN-CeNPs) integrates catalytic material design and immunomodulatory functions under the P–A–N framework. In the preparation phase, MSNs are engineered with high antigen-loading capacity to deliver self-antigens to splenic antigen-presenting cells (APCs), while surface-conjugate CeNPs provide catalytic ROS-scavenging activity via redox cycle-mediated antioxidant mimicry (e.g., SOD/CAT-like activity) (Fig. 3(d)). During activation, CeNPs catalytically neutralize inflammatory ROS in the microenvironment, synergizing with antigen-loaded MSNs to reprogram APCs into a tolerogenic phenotype marked by reduced costimulatory molecule (CD80/CD86) expression and anti-inflammatory cytokine secretion (Fig. 3(e)). This catalytic remodeling of APCs drives naïve T cell differentiation into Foxp3+ regulatory T cells (Tregs) rather than pathogenic Th1/Th17 subsets, establishing antigen-specific immune tolerance. In the nontoxic treatment phase, the expanded Treg population suppresses autoreactive CD4+ T cell infiltration into the CNS, halting demyelination and mitigating late-stage experimental autoimmune encephalomyelitis without systemic immunosuppression (Fig. 3(f)). By coupling catalytic material properties (ROS elimination) with biological immunomodulation (APC/Treg axis), this strategy exemplifies the transformative potential of pancatalytic biomaterials in resolving chronic neuroinflammation through spatially confined, toxicity-free therapeutic catalysis.51

3.3 Respiratory system

Inflammatory respiratory diseases represent a significant component of the global disease burden, with their development closely linked to multifaceted environmental and genetic risk factors.140 Exogenous stimuli such as environmental pollutants (e.g., PM2.5), cigarette smoke, and pathogenic microbial infections trigger chronic inflammatory responses, leading to progressive damage to airways and pulmonary parenchyma.141 Chronic inflammatory disorders, including asthma and chronic obstructive pulmonary disease (COPD), are characterized by high recurrence rates and disability, significantly impairing patients’ quality of life. Infectious diseases like bacterial pneumonia and tuberculosis face escalating therapeutic challenges due to the spread of drug-resistant strains and immune evasion mechanisms. Notably, fibrotic conditions such as idiopathic pulmonary fibrosis exhibit pathological progression intertwined with persistent inflammatory microenvironments, manifesting as inflammatory-fibrotic transitions.142

The shared pathological mechanisms of these diseases stem from dysregulated inflammatory response networks. PAMPs or DAMPs activate pattern recognition receptors (e.g., TLRs) on alveolar macrophages and airway epithelial cells, initiating NF-κB and MAPK signaling pathways.143 This cascade drives excessive release of pro-inflammatory cytokines (IL-1β, IL-6, TNF-α) and chemokines (C–X–C motif ligand 8, CXCL8; C–C motif ligand 2, CCL2).144,145 Recruitment of neutrophils and monocytes amplifies inflammatory cascades, with tissue destruction mediated by elastase and matrix metalloproteinases.146 Oxidative stress perpetuates a vicious cycle: ROS overproduction directly damages epithelial barriers and activates the NLRP3 inflammasome, promoting IL-1β maturation. Th2 immune polarization dominates allergic inflammation, where interleukin-4, interleukin-5, and interleukin-13 induce goblet cell hyperplasia and immunoglobulin E production and eosinophil infiltration, driving airway hyperresponsiveness.147 In chronic inflammation, dysfunctional Tregs and sustained activation of pro-inflammatory M1 macrophages jointly sustain pathological processes.148,149

COPD, ranked as the third leading global cause of mortality, imposes a substantial burden on public health through progressive airway obstruction, emphysema, and irreversible airflow limitation.148 Cigarette smoke remains the primary etiological driver, yet persistent airway inflammation persists even after smoking cessation, highlighting the complexity of its pathogenesis.150 COPD progression is critically mediated by dysregulated alveolar macrophage polarization in response to environmental triggers.145 M1 macrophages, activated by Th1 signals, exacerbate inflammation via overproduction of NO, TNF-α, IL-6, and CCL2, driving tissue destruction.144 Conversely, M2 macrophages, regulated by Th2 cytokines (IL-4/IL-13), promote anti-inflammatory responses and tissue repair through Arg1/TGF-β upregulation. However, the imbalance between M1-driven inflammation and M2-mediated repair mechanisms underpins chronic pathological changes, including airway remodeling and fibrosis.144 Current therapeutic strategies, such as inhaled corticosteroids and long-acting bronchodilators, offer symptomatic relief but fail to reverse structural damage or halt disease progression, particularly in advanced emphysema.151 These limitations underscore the urgent need for innovative therapeutic strategies capable of targeting the root causes of COPD pathology while overcoming anatomical and biochemical barriers in the respiratory tract. The PEG@CS/BPQDs-AM NPs exemplify a pancatalytic strategy for COPD therapy by integrating sequential catalytic processes into a unified therapeutic platform. This system addresses the challenges of mucus barrier penetration and controlled drug release through rationally engineered biomaterial interactions. The chitosan core, functionalized with polyethylene glycol (PEG), leverages its inherent mucoadhesive properties to adhere to the negatively charged pulmonary epithelium while enhancing mucus penetration via PEG-mediated hydrophilicity, a critical step in overcoming airway defenses. Embedded black phosphorus quantum dots (BPQDs) serve as catalytic triggers, where their gradual autoxidation generates phosphate ions and protons, inducing localized dissociation of the PEG@CS matrix. This self-regulated degradation not only ensures precise spatiotemporal release of anti-inflammatory macrophages at diseased sites but also disrupts pathogenic biofilms through reactive species generation, synergistically enhancing therapeutic efficacy. The pancatalytic framework is further realized by BPQDs’ inherent biocompatibility and biodegradability, ensuring nontoxic metabolic clearance post-treatment (Fig. 4(a)). By orchestrating material preparation (P), activation of the bio-effect (A), and nontoxic treatment (N), this system demonstrates holistic management of COPD pathology, offering a paradigm for designing multifunctional biomaterials that harmonize catalytic processes with biological precision.52


image file: d5nh00408j-f4.tif
Fig. 4 Pancatalytic biomaterials for the treatment of respiratory system diseases. (a) PEG@CS/BPQDs-AM NPs for COPD treatment via the P–A–N framework. Reproduced with permission.52 Copyright 2020 Wiley-VCH GmbH. (b) CAT-NGs for asthma treatment via the P–A–N framework. Reproduced with permission.40 Copyright 2024 Wiley-VCH GmbH.

Asthma, a chronic respiratory disease marked by airway inflammation, bronchoconstriction, and mucus overproduction, affects over 300 million people globally, with rising prevalence linked to genetic-environmental interactions.152,153 Neutrophilic asthma, a severe subtype resistant to corticosteroids and associated with bacterial colonization (e.g., NTHi, S. aureus), represents 5–20% of cases but accounts for 50–80% of asthma-related healthcare burden due to high hospitalization and mortality.154,155 Pathological mechanisms involve oxidative stress from excessive ROS generated by pollutants, allergens, and inflammatory cells via Nox2-derived superoxide, which transforms into damaging H2O2 and downstream radicals (e.g., ˙OH, HOCl), worsening airway hyperresponsiveness and remodeling.156–158 While CAT demonstrates potent H2O2-neutralizing capacity, its clinical utility is hampered by rapid degradation and poor bioavailability.159 Current therapies, including corticosteroids and bronchodilators, inadequately address ROS-driven pathology, particularly in neutrophilic variants, necessitating innovative strategies to enhance antioxidant delivery and bacterial clearance.160 The CAT-loaded nanogels (CAT-NGs) embody a pancatalytic approach by integrating catalytic ROS scavenging with multifunctional biomaterial engineering. Composed of chitosan–arginine–maleimide conjugates crosslinked with citric acid and ε-polylysine, these nanogels protect CAT from enzymatic degradation while leveraging their mucoadhesive properties for targeted airway delivery. The system operates through sequential catalytic processes, where ionic crosslinking during preparation ensures structural stability, while the antibacterial ε-polylysine shell activates biofilm disruption to counteract infection-associated neutrophilic inflammation (Fig. 4(b)). Upon inhalation, CAT-NGs neutralize H2O2 via enzymatic catalysis, interrupting ROS propagation and suppressing NLRP3/NF-κB signaling pathways, a nontoxic intervention that concurrently alleviates oxidative stress and inflammation. This holistic strategy not only preserves CAT's catalytic efficiency but also addresses comorbid bacterial challenges, demonstrating how pancatalytic biomaterials synchronize material design, biological activation, and therapeutic precision to redefine asthma management.40

3.4 Digestive system

Inflammatory gastrointestinal diseases represent a heterogeneous group of disorders characterized by chronic inflammation of the digestive tract, severely compromising patients’ quality of life and long-term health outcomes.161 Among these, IBD and metabolic dysfunction-associated steatohepatitis (MASH) stand as two prototypical pathological entities.162 IBD patients frequently suffer from persistent intestinal inflammation, leading to chronic diarrhea, malabsorption, progressive malnutrition, weight loss, and growth retardation in pediatric populations. Prolonged mucosal damage significantly elevates the risk of colorectal cancer, with severe cases necessitating surgical resection of affected bowel segments.163,164 MASH, on the other hand, is defined by metabolic dysregulation-driven hepatic inflammation and fibrosis, placing patients at heightened risk of end-stage liver diseases such as cirrhosis, hepatic decompensation, and hepatocellular carcinoma.165 Both conditions are further complicated by extraintestinal systemic manifestations, including arthritis, uveitis, and erythema nodosum, which amplify patients’ physical and psychological burden.166

Inflammatory bowel diseases (IBDs), encompassing Crohn's disease (CD) and ulcerative colitis (UC), represent chronic and relapsing inflammatory disorders of the gastrointestinal tract, imposing substantial burden on global healthcare systems and patient well-being.167 Characterized by heterogeneous pathology, CD manifests as transmural, discontinuous inflammation across the entire gastrointestinal tract, predominantly targeting the ileum and colon, while UC is restricted to continuous mucosal inflammation in the colon and rectum.168,169 Both subtypes severely compromise patients’ quality of life, with elevated risks of colorectal cancer, systemic complications, and mortality, underscoring their clinical urgency.170 The global prevalence of IBD has surged, affecting over 3.5 million individuals globally, driven by complex interactions among genetic susceptibility, environmental triggers (e.g., dietary shifts, psychological stress), and microbial dysbiosis.171–174 Despite advances in identifying these contributing factors, the precise etiology remains elusive, hindering the development of curative therapies.

Current management strategies predominantly rely on immunosuppressive agents, corticosteroids, and biologics such as anti-TNF monoclonal antibodies, which aim to attenuate aberrant immune activation and promote mucosal healing.175,176 However, therapeutic efficacy is inconsistent, with up to 30% of patients exhibiting primary non-response to anti-TNF agents and 30–40% developing secondary resistance over time, necessitating escalated dosing or alternative regimens, which amplify the risk of infections and malignancies.177–180 Surgical interventions, though employed in severe cases, are limited by prolonged recovery, recurrence, and functional impairments. Conventional drug delivery systems, including pH-responsive or enzyme-triggered formulations, face challenges in overcoming the colon's hostile microenvironment, marked by fluctuating pH, enzymatic degradation, mucus barriers, and microbial activity, leading to premature drug release or inadequate targeting.181 These limitations highlight the unmet need for advanced biomaterial-based platforms capable of precise, sustained drug delivery to inflamed tissues while minimizing systemic toxicity.

To address these critical gaps, RuCo nanosheets have emerged as a promising pancatalytic approach for IBD treatment through integrated preparation, activation, and nontoxic action. In the preparation phase, bimetallic RuCo nanosheets are synthesized via solvothermal methods, leveraging ruthenium's acid resistance and cobalt's biocompatibility to optimize stability and catalytic efficiency. Their electronic structure modulation, driven by electronegativity differences, enhances dual enzyme-like activities (SOD and CAT), validated by density functional theory calculations. During activation, RuCo nanosheets catalytically scavenge ROS via SOD-like and CAT-like cascades, disrupting NF-κB signaling and suppressing proinflammatory cytokine release. This mitigates intestinal barrier damage and dysbiosis, addressing IBD's core pathology. The nontoxic treatment phase is ensured by cobalt's physiological relevance and ruthenium's gastric stability, enabling oral delivery with minimal systemic exposure (Fig. 5(a)). Preclinical studies demonstrate significant anti-inflammatory efficacy without adverse effects, highlighting targeted ROS neutralization and sustained catalytic activity in the gastrointestinal tract. By unifying material design, catalytic bioactivity, and biocompatibility, RuCo nanosheets embody pancatalysis, offering a robust, orally administrable strategy for IBD management.53


image file: d5nh00408j-f5.tif
Fig. 5 Pancatalytic biomaterials for the treatment of digestive system diseases. (a) RuCo nanosheets for IBD treatment via the P–A–N framework. Reproduced with permission.53 Copyright 2025 American Chemical Society. (b) LNPs function as pancatalytic biomaterials in metabolic dysfunction-associated MASH. Reproduced with permission.54 Copyright 2024 Springer Nature.

MASH, a severe subtype of metabolic dysfunction-associated steatotic liver disease, has emerged as a global health crisis driven by chronic caloric excess, genetic susceptibility, and metabolic dysregulation.182 Characterized by hepatic steatosis, lobular inflammation, hepatocyte injury, and progressive fibrosis, MASH significantly elevates the risk of cirrhosis, hepatocellular carcinoma, and liver-related mortality, while imposing a growing burden on healthcare systems worldwide. Its pathophysiology involves a complex interplay of lipotoxicity, mitochondrial dysfunction, oxidative stress, and dysregulated immune responses.182 Lipid nanoparticles (LNPs) function as pancatalytic biomaterials in MASH therapy through a holistic strategy integrating preparation, activation, and nontoxic targeting (Fig. 5(b)). The preparation phase involves engineering LNPs with retinoid derivatives (e.g., all-trans retinoic acid) to enhance fibrotic liver targeting. These ligands bind retinol-binding protein-4 (RBP-4) on activated hepatic stellate cells (HSCs), with carboxylic retinoids optimizing LNP surface rearrangement for improved RBP-4 affinity and cellular uptake. This design achieves a ∼10-fold increase in mRNA delivery efficacy compared to conventional LNPs. In the activation phase, LNPs deliver mRNA encoding therapeutic proteins like relaxin, an anti-fibrotic agent targeting the TGF-β pathway, fused to a collagen-binding domain (CBD) for exopolysaccharide-rich extracellular matrix (ECM) anchoring. This fusion enables ∼80% hepatic retention, ensuring sustained local activity to suppress HSC activation and ECM deposition. Finally, the nontoxic treatment phase minimizes systemic exposure by combining RBP-4-mediated targeting and CBD-driven ECM anchoring, reducing off-target effects while maintaining therapeutic efficacy. Validated in preclinical MASH models, this pancatalytic approach exemplifies precise catalyst design, bioactivity amplification, and nontoxic treatment, offering a transformative framework for fibrosis treatment.54

3.5 Joint systems

Osteoarthritis (OA) and rheumatoid arthritis (RA), two prevalent forms of chronic inflammatory arthritis, collectively impose substantial burden on global health due to their progressive destruction of synovial joints and irreversible functional impairment. While OA primarily stems from mechanical wear-and-tear of articular cartilage, RA arises from autoimmune-driven synovial tissue attacks, though both culminate in a vicious cycle of joint degradation and disability.183 The pathological difference lies in their destructive patterns: RA predominantly induces synovial hyperplasia and superficial osteochondral erosion through autoimmune inflammation, whereas OA triggers extensive structural breakdown penetrating deep into subchondral bone layers.72 Shared between these disorders is the establishment of a chronically inflamed microenvironment characterized by hyperactivated immune cell infiltration and sustained pro-inflammatory cytokine storms, collectively termed the “inflammatory arthritic cascade”.184 Clinically, this manifests as progressive joint space narrowing, bony ankylosis, and eventual deformity with mobility loss. Current therapeutic regimens, including anti-inflammatories, disease-modifying anti-rheumatic drugs (DMARDs), and immunosuppressants, merely alleviate symptoms while failing to halt disease progression, often causing side effects upon prolonged administration.185 These limitations underscore the urgent need for targeted therapies capable of resolving localized inflammation without compromising systemic immunity, a challenge being addressed by emerging pancatalytic biomaterial strategies.

RA, a chronic autoimmune disorder affecting ≈0.5–1.0% of the global population, leads to debilitating joint inflammation, bone erosion, and progressive deformity, severely compromising patients’ quality of life.186,187 Its pathogenesis involves misguided immune attacks on synovial tissues, triggering a self-perpetuating cycle of oxidative stress and inflammatory cascades.187 Excessive ROS not only damage bone and cartilage but also polarize macrophages toward pro-inflammatory M1 phenotypes, amplifying cytokine storms (e.g., IL-6, TNF-α) that exacerbate the inflammatory cascades. Current clinical strategies, including NSAIDs, glucocorticoids, DMARDs like methotrexate, and biologics such as TNF-α/IL-6 inhibitors, face critical limitations. While these agents temporarily suppress inflammation, they often cause systemic toxicities (e.g., gastrointestinal bleeding, osteoporosis) or exhibit inadequate therapeutic durability.188,189 High costs and variable patient responses further restrict the accessibility and efficacy of advanced biologics. A central challenge lies in reconciling targeted immunomodulation with tissue regeneration, as conventional therapies fail to address both inflammatory dysregulation and structural joint repair.

The pancatalytic biomaterial platform, exemplified by calcium phosphate particles (CAPs) derived from hydrolyzed CSNs, offers a holistic solution by integrating catalytic and quasi-catalytic processes. In the preparation (P) phase, CSNs are engineered through liquid exfoliation of bulk CaSi2, ensuring controlled hydrolysis kinetics (Fig. 6(a)-i and ii). Upon reaching inflamed joints, CSNs undergo activation (A) via hydrolysis, releasing hydrogen gas (H2) to scavenge ROS, alkaline Ca(OH)2 to inhibit osteoclast activity, and Ca2+/SiO2 precursors for CAP formation. The catalytic H2 neutralizes oxidative stress while modulating macrophage polarization, breaking the ROS–inflammation feedback loop (Fig. 6(b)-i–iii). Simultaneously, localized alkalinity disrupts osteoclastic bone resorption, and endogenous Ca2+/phosphate interactions drive the nontoxic treatment (N) phase by mineralizing CAPs, which promote osteogenic repair without exogenous toxicity (Fig. 6(c)). This tripartite pancatalytic strategy, synchronizing antioxidant, anti-inflammatory, and bone-remodeling actions, addresses RA's multifaceted pathology. By leveraging self-sustaining catalytic cascades and biocompatible byproducts, the system circumvents systemic side effects while achieving sustained therapeutic outcomes, embodying the pancatalysis principle of holistic disease management through material-catalyzed biological modulation.25


image file: d5nh00408j-f6.tif
Fig. 6 Pancatalytic biomaterials for the treatment of joint system diseases. (a-i) Schematic representation of CSNs fabricated from bulk calcium disilicide (BCS) through ultrasonic-assisted exfoliation. (a-ii) TEM image of CSNs with uniform size. Scale bar: 500 nm. (b-i) Schematic representation of the reaction product from the hydrolysis of CSNs. (b-ii) Photograph and quantitative measurements of hydrogen gas produced by CSNs reacting with H2O at different pH values. (b-iii) Electron spin resonance analysis of CSNs in the presence of Fe2+ (20 μm) and H2O2 (2.5 mm). (c-i) Representative photographs of the ankle joint of mice from different groups. (c-ii) body weight curves of mice subject to different treatments. Data are expressed as means ± SD (N = 5). (c-iii) Representative 3D-reconstructed micro-CT images of right ankle joints of mice from different groups on day 27. Histomorphometric micro-CT analysis of bone microstructure. (c-iv) Tb.BMD and (c-v) Tb.Sp for right ankle joints of mice from different groups on day 27. Data are expressed as means ± SD (N = 4). Reproduced with permission.25 Copyright 2024 Wiley-VCH GmbH. (d) BPNSs for OA treatment via the P–A–N framework. (e-i) Synthesis of BPNSs through a modified sonication-assisted liquid exfoliation method. Representative (e-ii) SEM and (e-iii) TEM images of BPNSs. The scavenging activity of BPNSs toward three typical free radicals was monitored via UV-vis spectroscopy and quantitatively analyzed: (f-i) DPPH and (f-ii) ABTS. (g) Representative three-dimensional reconstruction images of differently treated rat knees at weeks 4 and 8 (upper row) and coronal micro-CT images of the rat medial tibial plateau in different groups (lower row). Reproduced with permission.37 Copyright 2023 American Chemical Society.

OA, a prevalent degenerative joint disorder affecting over 52 million adults in the U.S. alone, imposes significant socioeconomic burden due to its association with chronic pain, cartilage destruction, and progressive joint dysfunction.190 While traditionally viewed as a cartilage-centric disease, OA is now recognized as a whole-joint pathology involving synovial inflammation, subchondral bone remodeling, and osteophyte formation.191 Central to its progression is a self-amplifying cycle of oxidative stress and mitochondrial dysfunction in chondrocytes, where defective mitochondria generate excessive ROS that exacerbate cartilage degradation, ATP depletion, and apoptotic cascades.192–194 Subchondral bone alterations further complicate OA pathogenesis, shifting from early-stage bone resorption to late-stage sclerosis, with evidence suggesting that restoring subchondral bone architecture could potentially reverse disease progression.195–197 While non-surgical approaches like medication and physical therapy can alleviate pain and preserve joint function to improve quality of life, they remain limited to symptom control and fail to target core pathological processes such as cartilage degeneration and synovial inflammation. More importantly, current therapies cannot effectively halt disease progression or achieve tissue repair, and this fundamental limitation of “addressing symptoms without resolving the root cause” significantly hinders the overall efficacy of OA treatment.198

The pancatalytic biomaterial approach, exemplified by BPNSs, addresses OA's multifaceted pathology through integrated catalytic processes (Fig. 6(d)). In the preparation (P) phase, BPNSs are engineered via liquid-phase exfoliation of bulk black phosphorus, leveraging weak van der Waals interactions to create atomically thin 2D structures with inherent biocompatibility (Fig. 6(e)). Upon reaching OA-affected joints, BPNSs undergo activation (A) through surface oxidation, where their unique electronic structure enables catalytic ROS scavenging via electron transfer reactions, neutralizing free radicals while suppressing pro-inflammatory macrophage polarization (Fig. 6(f)). This antioxidant activity disrupts the ROS-mitochondrial dysfunction cycle, preserving chondrocyte viability and cartilage homeostasis. Simultaneously, BPNSs’ degradation products, primarily phosphorus oxides, initiate nontoxic treatment (N) by promoting subchondral bone regeneration through phosphate ion release, which stimulates osteogenic mineralization without exogenous toxicity (Fig. 6(g)). This dual-action mechanism concurrently targets oxidative stress in cartilage and bone remodeling in subchondral regions, embodying the pancatalysis principle of harmonizing catalytic bioactivity (ROS elimination) with quasi-catalytic biological processes (bone regeneration). By coupling material-driven catalytic ROS neutralization with physiologically guided tissue repair, BPNSs achieve holistic OA management that surpasses conventional symptom-focused therapies, demonstrating how pancatalytic systems can synchronize disease-modifying interventions across multiple pathological axes.37

3.6 Integumentary system

Skin inflammatory diseases affect a significant portion of the population, with estimates suggesting that 20–25% of the population are impacted by chronic, non-communicable forms.199 Specific conditions show varying prevalence: atopic dermatitis (AD) affects up to 30% of children and 2–10% of adults, while psoriasis impacts 2–3% of the population.200 Hidradenitis suppurativa ranges from 0.2 to 0.4% on average, with higher rates in certain demographics like 1.3% in African Americans.199 These diseases significantly impair quality of life, with AD linked to sleep disturbances and mental health issues like depression and suicide risk, and psoriasis associated with cardiovascular comorbidities, reducing life expectancy by about 15 years in severe cases.201

AD represents a chronic inflammatory skin disorder marked by persistent pruritus, epidermal barrier dysfunction, and immune dysregulation, imposing substantial physical, psychological, and socioeconomic burden globally.202–204 Predominantly emerging in early childhood but increasingly prevalent in adults, AD significantly compromises quality of life for patients and caregivers while elevating the risk of comorbidities such as allergic diseases, cardiovascular disorders, and mental health conditions.205–209 Its pathogenesis involves multifaceted interactions among genetic predisposition, epidermal barrier impairment, immune hyperactivation (particularly Th2-driven inflammation), and oxidative stress.210 Chronic inflammation in AD exacerbates ROS accumulation, which further disrupts barrier integrity, perpetuates pruritus through the itch–scratch cycle, and amplifies inflammatory cascades.211 Despite current therapies targeting anti-inflammatory pathways, allergen avoidance, and phototherapy, limitations persist, including incomplete efficacy, systemic side effects, and the need for prolonged monitoring, especially in pediatric populations. These challenges underscore the demand for novel strategies that address the root mechanisms of AD, particularly oxidative-antioxidant imbalances and immune dysregulation.212 In this context, pancatalytic biomaterials like the copper/zinc metal–organic framework (Cu/Zn-MOF) offer a holistic therapeutic approach by integrating catalytic preparation (P), activation (A) of the bio-effect, and nontoxic treatment (N). Cu/Zn-MOF mimics the dual-site synergistic action of natural antioxidant enzymes such as SOD and GPx, which are deficient in AD patients. By replicating the histidine-bridged Cu–Zn bimetallic coordination found in CuZn-SOD, this biomaterial scavenges ROS through catalytic dismutation of superoxide radicals, thereby restoring redox homeostasis (Fig. 7(a) and (b)). Furthermore, Cu/Zn-MOF suppresses Fcγ receptor-mediated phagocytic signaling, a pathway implicated in immune complex-driven inflammation, thereby mitigating chronic inflammatory responses in AD lesions (Fig. 7(c)). The pancatalytic strategy not only addresses oxidative stress but also modulates immune hyperactivity without relying on conventional immunosuppressants, aligning with the need for safer, multitarget therapies. By bridging catalytic material design with disease-specific mechanisms, this approach exemplifies the translation of biomimetic catalysis into clinical applications, offering a paradigm for managing complex inflammatory skin disorders through redox regulation and immune pathway modulation.55


image file: d5nh00408j-f7.tif
Fig. 7 Pancatalytic biomaterials for the treatment of integumentary system diseases. (a-i) Schematic illustration of Cu/Zn-MOF synthesis. (a-ii) TEM images of Cu/Zn-MOF. (a-iii) Element mapping displaying the distribution of C, N, Zn, and Cu. (a-iv) AC-HAADF-STEM image of Cu/Zn-MOF. (b-i) Schematic illustration of the ROS cascade clearance of Cu/Zn-MOF. (b-ii) SOD-like performance of Cu/Zn-MOF under various concentrations. (b-iii) GPx-like activity evaluation of Cu/Zn-MOF. (iv) TPx-like activity analysis by Cu/Zn-MOF concentration-dependent cysteine (0.2 mM) depletion in the presence of H2O2. (c-i) Images of skin lesions after 14-day induction (n = 4 samples/group). (c-ii) Quantification of ear thickness. Reproduced with permission.55 Copyright 2024 American Chemical Society. (d-i) Synthetic process of biomimetic FeN4O2-SACs. (d-ii) TEM image of FeN4O2-SACs. (d-iii) AC HAADF-STEM image of FeN4O2-SACs, with single iron atoms marked using red circles. n = 3 samples with similar results. (d-iv) HAADF-STEM image and the corresponding EDX mappings of FeN4O2-SACs: C, blue; N, green; Fe, yellow. (d-v) EELS atomic spectrum of Fe elements in FeN4O2-SACs. The element-specific absorption edge is highlighted by a red circle. (e-i) Schematic illustration of ROS clearance by FeN4O2-SACs. (e-ii) ESR spectra of FeN4O2-SACs and natural SOD for O2˙ clearance at pH = 7.4 (X: xanthine; XO: xanthine oxidase). (e-iii) Comparison of SOD-like activities between FeN4O2-SACs and control materials CeO2, Mn3O4, Co-SACs, Cu-SACs and Zn-SACs by WST-8 colorimetric analysis. (e-iv) CAT-like assay for eliminating varied concentrations of H2O2 using 125[thin space (1/6-em)]μg ml−1 FeN4O2-SACs by UV absorption tests. (f-i) Representative images of lesions on day 8, n = 4 samples per group. (f-ii) Quantification of ear thickness. Reproduced with permission.56 Copyright 2023, Springer Nature.

Psoriasis is a chronic inflammatory skin disorder characterized by recurrent erythematous plaques, abnormal keratinocyte proliferation, and dysregulated immune responses, affecting approximately 3% of the global population.213–216 Its high recurrence rate (approximately 90%) and association with systemic comorbidities, including CVDs and diabetes, impose substantial physical, psychological, and socioeconomic burden.217 Pathologically, excessive ROS accumulation driven by oxidative-antioxidative imbalance amplifies inflammatory cytokine release, perpetuates keratinocyte hyperproliferation, and sustains chronic inflammation.218–220 Current therapies, such as immunosuppressants and biologics, often fail to address root causes, leading to relapse post-treatment, increased infection risks, and potential carcinogenicity. These limitations highlight the urgent need for therapies that target oxidative stress and immune dysregulation without compromising long-term safety.221–223 The pancatalytic approach using biomimetic iron single-atom catalysts (FeN4O2-SACs) offers a holistic strategy by integrating catalyst preparation (P), activation (A) of the bio-effect, and nontoxic treatment (N) of diseases. FeN4O2-SACs mimic the catalytic centers of multiple natural antioxidant enzymes, including SOD, CAT, and ascorbate peroxidase, through their atomically dispersed FeN4O2 coordination structure. This design enables broad-spectrum ROS scavenging via catalytic dismutation of ˙O2 and H2O2, restoring redox homeostasis in psoriatic lesions (Fig. 7(d) and (e)). By suppressing ROS-driven inflammatory cascades and keratinocyte hyperproliferation, FeN4O2-SACs alleviate symptoms and reduce relapse rates, outperforming conventional treatments like calcipotriol (Fig. 7(f)). Mechanistically, FeN4O2-SACs upregulate estrogen receptor 1, a key protein downregulated in psoriasis, thereby modulating inflammatory pathways and enhancing therapeutic durability. The pancatalytic framework not only addresses oxidative stress through enzyme-like catalysis but also ensures biocompatibility and sustained efficacy, exemplifying a multitarget, non-immunosuppressive strategy for managing psoriasis. This biomimetic approach bridges atomic-level catalyst design with disease-specific mechanisms, advancing catalytic therapy as a paradigm for inflammatory skin disorders.56

3.7 Sensory systems

The sensory system serves as the body's primary network for environmental perception, integrating organs such as the eyes, ears, nose, tongue, and skin to mediate vision, hearing, smell, taste, and tactile functions.224 Ocular and auditory systems are particularly critical for information processing and quality of life; inflammatory disorders in these systems, including conjunctivitis (conjunctival inflammation), keratitis (corneal inflammation), uveitis (uveal tract inflammation), and dry eye syndrome, often induce symptoms such as chronic irritation, photophobia, and progressive vision deterioration.225 In the auditory domain, SNHL, frequently associated with labyrinthitis (inner ear inflammation), manifests as sudden hearing impairment, tinnitus, and balance disturbances.226 Beyond direct sensory dysfunction, these conditions exert profound systemic impacts: visual deficits may restrict social engagement and exacerbate psychological distress, while auditory impairments amplify communication barriers, fostering social isolation and depression.227 Chronic inflammation in untreated cases precipitates irreversible complications, including corneal scarring, secondary glaucoma, and persistent vestibular dysfunction, severely compromising occupational productivity and functional independence.228 Therapeutic strategies emphasize early intervention with anti-inflammatory agents or surgical approaches to mitigate long-term sequelae, highlighting the urgency of targeted therapies to address sensory system inflammations in clinical practice.

SNHL, a major global public health issue, severely compromises social communication and imposes significant socioeconomic burden due to irreversible hair cell (HC) loss and spiral ganglion neuron damage caused by factors such as noise exposure, genetic mutations, aging, and ototoxic drugs like cisplatin.229–231 Cisplatin-induced ototoxicity is characterized by progressive and irreversible hearing impairment, primarily driven by oxidative stress-triggered inflammatory cascades and metabolic disturbances, while current clinical interventions (e.g., hearing aids or cochlear implants) are limited by invasiveness, high costs, and inability to repair cellular damage.232–234 To address these challenges, the Zr-MOF-Mn composite, guided by the pancatalysis strategy (P–A–N), offers a synergistic therapeutic approach Fig. 8(a). The Zr-MOF-framework, engineered with tunable mesopores, serves as a carrier to integrate Mn-TCPP (a superoxide dismutase-mimetic enzyme) and N-acetyl-L-cysteine-derived carbonized polymer dots (NAC CPDs), leveraging its high surface area and biocompatibility for efficient enzyme delivery. Mn-TCPP scavenges ROS via Mn3+/Mn4+ redox cycling, while NAC CPDs enhance chemical stability and prolong antioxidant efficacy. Upon crossing the round window membrane into the cochlea, the composite targets HC mitochondria to eliminate ROS, thereby suppressing pathological HIF-1α accumulation, reversing glycolysis-dominated metabolic reprogramming, and restoring oxidative phosphorylation. Simultaneously, it inhibits BAX-mediated mitochondrial outer membrane permeabilization and mtDNA leakage, blocking caspase-dependent apoptosis and NLRP3 inflammasome activation. To ensure sustained therapeutic action and minimize systemic toxicity, PDA-coated GelMA hydrogel microcapsules are employed, which enable controlled drug release, extending local antioxidant effects within the cochlea. This pancatalysis-driven system achieves mitochondrial homeostasis restoration and immune microenvironment modulation, providing a multifunctional “ROS scavenging–apoptosis inhibition–anti-inflammation” paradigm that overcomes the limitations of conventional therapies for cisplatin-induced SNHL.57

Dry eye disease (DED), a prevalent global ocular surface disorder affecting approximately one-third of the population, manifests as chronic ocular discomfort, visual impairment, and diminished quality of life due to tear film instability, sustained inflammation, and corneal microenvironment imbalance.235,236 Its pathogenesis involves complex interactions among intrinsic and extrinsic factors, including oxidative stress–inflammation vicious cycles, where excessive ROS amplify inflammatory signaling, exacerbating epithelial damage and immune dysregulation.237–240 Conventional therapies, such as artificial tears and immunosuppressants (e.g., cyclosporine A), provide only transient symptom relief and face limitations like low bioavailability, local irritation, and an inability to simultaneously target oxidative stress and inflammatory pathways.241 Addressing these challenges, the Cs@P/CeO2 composite, guided by the pancatalysis strategy (P–A–N), offers a synergistic therapeutic approach (Fig. 8(b)). The CeO2 nanozyme core, engineered with oxygen vacancies enabling Ce3+/Ce4+ redox cycling, provides self-regenerative antioxidative capacity. PEG surface modification enhances biocompatibility and hydrophilicity, while cyclosporine A (CsA) loading integrates immunomodulatory functionality. Upon ocular administration, Cs@P/CeO2 penetrates the ocular surface to scavenge ROS, disrupt the oxidative stress–inflammation cascade, and suppress T-cell-mediated hyperimmunity. Single-cell sequencing reveals its capacity to reprogram the corneal microenvironment by polarizing macrophages from pro-inflammatory (M1) to pro-repair (M2) phenotypes, promoting epithelial regeneration, and fine-tuning immune–epithelial crosstalk for tissue homeostasis. The PEG shell minimizes systemic toxicity and extends corneal retention, enabling sustained drug release, while the self-renewing antioxidative activity of nanoceria ensures long-term efficacy. By holistically integrating catalytic ROS scavenging, immunomodulation, and biocompatible delivery, Cs@P/CeO2 exemplifies the pancatalysis-driven paradigm, overcoming the limitations of single-target therapies and offering a “ROS elimination–inhibition of inflammation–tissue repair” strategy for DED and related oxidative stress–inflammatory disorders.45


image file: d5nh00408j-f8.tif
Fig. 8 Pancatalytic biomaterials for the treatment of sensory system diseases. (a) Treatment of SHLH based on Zr-MOF-Mn composite materials via the P–A–N framework. Reproduced with permission.57 Copyright 2024 Wiley-VCH GmbH. (b) Cs@P/CeO2 for DED treatment via the P–A–N framework. Reproduced with permission.45 Copyright 2024 American Chemical Society. (c) DCFH system for uveitis treatment via the P–A–N framework. Reproduced with permission.58 Copyright 2023 American Chemical Society. (d) MLPGa nanosystem for fungal keratitis treatment via the P–A–N framework. Reproduced with permission.59 Copyright 2022 Wiley-VCH GmbH.

Uveitis, a spectrum of immune-mediated ocular disorders affecting the uveal tract and adjacent tissues, manifests as ocular pain, redness, photophobia, and vision-threatening complications, contributing to 10–20% of blindness cases in the United States due to recurrent inflammation and secondary damage.242,243 Its pathogenesis involves dysregulated immune responses intertwined with oxidative stress, where excessive ROS and NO overproduction disrupt the blood–aqueous barrier, induce lipid peroxidation in iris epithelial cells and photoreceptor mitochondria, and amplify inflammatory cascades.244–246 Current first-line therapies, such as corticosteroids (e.g., dexamethasone sodium phosphate, DSP), face limitations including incomplete efficacy (e.g., 25% non-response rate in clinical trials) and adverse effects like intraocular pressure elevation.247,248 Moreover, conventional drug delivery routes, noninvasive eye drops with low bioavailability or invasive methods risking infection, fail to ensure sustained therapeutic concentrations. To address these challenges, the DSP@Ce-MOFs F127 hydrogel (DCFH) system, guided by the pancatalysis strategy (P–A–N), integrates multifunctional biomaterials for synergistic therapy (Fig. 8(c)). Preparation (P): cerium-based metal–organic frameworks (Ce-MOFs) are engineered as porous carriers with dual functionality, ROS scavenging via Ce3+/Ce4+ redox cycling and high-capacity DSP loading due to their large surface area. Activation (A): Ce-MOFs demonstrate dual functionality through their ROS-scavenging capacity and porous structure that enables efficient DSP loading/release, while the F127 hydrogel enhances ocular retention and sustained drug delivery. Nontoxic treatment (N): the DCFH system achieves therapeutic efficacy by synergistically reducing oxidative stress and inflammation, while its ocular biocompatibility ensures safe application without adverse effects.58

Fungal keratitis, a leading cause of corneal blindness in developing countries, is characterized by severe infections (e.g., Candida albicans) that induce corneal opacity, pain, hypopyon, and irreversible vision loss.249,250 Its pathogenesis involves fungal biofilm formation and exopolysaccharide-rich extracellular matrix (ECM), particularly β-1,3-glucan, which act as permeation barriers to shield pathogens from antifungal agents.251,252 Biofilms exhibit innate resistance, requiring drug concentrations up to 1000-fold higher than those effective against planktonic fungi, rendering conventional topical antifungals (e.g., azoles) inadequate due to poor penetration and limited efficacy. Furthermore, the biofilm-associated ECM promotes fungal adherence and immune evasion, exacerbating treatment challenges.253,254 Current therapies fail to address both biofilm disruption and metabolic interference, highlighting the need for multifunctional strategies that degrade exopolysaccharides while enhancing antifungal potency. Guided by the pancatalysis framework (P–A–N), the lyticase and gallium ions co-integrated nanosystem (MLPGa) integrates catalytic degradation of fungal barriers and metal-ion-mediated antifungal activity (Fig. 8(d)). Preparation (P): the system employs PDA-modified mesoporous silica nanoparticles as carriers, co-loading lyticase (a β-1,3-glucanase) and gallium ions (Ga3+). Lyticase enzymatically degrades β-1,3-glucan in fungal cell walls and biofilms, while Ga3+ mimics iron to disrupt fungal metabolic pathways. Activation (A): upon application, lyticase catalytically hydrolyzes exopolysaccharides, dismantling biofilm integrity and exposing embedded fungi. Concurrently, Ga3+ penetrates the compromised fungal cells, acting as a “Trojan horse” to interfere with iron-dependent enzymes, induce metabolic dysfunction, and generate cytotoxic ROS. This dual action synergistically eliminates both planktonic fungi and resilient biofilms. Nontoxic treatment (N): the PDA coating enhances biocompatibility and enables real-time Raman monitoring of Ga3+ release, ensuring controlled ion delivery to minimize off-target toxicity. By combining enzymatic biofilm disruption, Ga3+-mediated metabolic interference, and biocompatible monitoring, MLPGa exemplifies a pancatalysis approach, holistically addressing fungal resistance mechanisms while prioritizing safety. This strategy overcomes the limitations of single-target antifungals, offering a “barrier degradation–metabolic disruption–toxicity mitigation” paradigm for fungal keratitis and other biofilm-associated infections.59

3.8 Urinary system

The urinary system is highly susceptible to a variety of IDs, including acute kidney injury (AKI) and urinary tract infections (UTIs).255,256 These conditions are not only among the most prevalent and recurrent clinical problems, but they also carry a profound risk of progressing to chronic kidney disease, organ dysfunction, or severe systemic complications. The complexity of these diseases lies in their multifactorial pathogenesis, including ischemic and oxidative insults or persistent bacterial colonization and inflammation.257 The high morbidity and mortality associated with these diseases, together with their tendency to cause chronic disability, underscore the urgency of developing more effective and targeted therapeutic strategies.

For AKI, which is characterized by a sudden decline in renal function and a high risk of progression to chronic kidney disease, the major pathophysiological trigger is renal ischemia/reperfusion injury, leading to excessive generation of ROS and mitochondrial dysfunction in renal proximal tubule cells.258,259 Despite the protective role of irisin, a myokine that promotes mitochondrial biogenesis and oxidative metabolism, its clinical application is severely limited by instability in circulation, short half-life, and poor kidney targeting.260 The Pt5.65S pre-nanozyme demonstrates pancatalytic biomaterial properties for AKI treatment by implementing the P–A–N framework through its unique design and multimodal therapeutic actions (Fig. 9(a)). Engineered via Ptzyme–Na2S reaction during the preparation phase, this ultra-small nanoparticle remains catalytically inert under normal physiological conditions. The activation mechanism involves two distinct but complementary therapeutic pathways (Fig. 9(b)-i and ii). First, the system releases H2S gas which mediates critical biological effects through sulfation of NF-κB/Iκκβ, effectively suppressing inflammatory signaling cascades while concurrently enhancing Nrf2-mediated antioxidant pathways. Second, the nanoparticle exhibits pH-triggered broad-spectrum ROS/RNS scavenging capacity by demonstrating both SOD-like and CAT-like enzymatic activities specifically in the acidic/inflammatory AKI microenvironment (Fig. 9(b)-iii). The system ensures nontoxic treatment through renal-clearable dimensions and conditional activation that prevents off-target effects, while combining exogenous nanozyme catalysis with endogenous H2S signaling for synergistic oxidative stress and inflammation regulation (Fig. 9(c)).60 This integrated approach exemplifies pancatalysis by simultaneously addressing multiple pathological factors (oxidative stress, inflammation, and cellular damage) through coordinated material design, microenvironment-responsive activation, and biosafety features tailored for AKI management.


image file: d5nh00408j-f9.tif
Fig. 9 Pancatalytic biomaterials for the treatment of urinary system diseases. (a) Pt5.65S pre-nanozyme for AKI via the P–A–N framework. (b-i) TEM image of the Pt5.65S pre-nanozyme (scale bar = 5 nm). (b-ii) Schematic illustration of Pt5.65S pre-nanozyme pH active antioxidative activities and release of H2S. (c-i) TEM images of mitochondria in the kidney from each group. The red box in the figure depicts representative TEM images showing the morphology of mitochondria under different conditions. Scale bar = 500 μm. (c-ii) HE staining of the heart, liver, spleen, lung, and kidney. Scale bar = 100 μm. Reproduced with permission.60 Copyright 2024 Advanced Science published by Wiley-VCH GmbH. (d) DEC for UTIs via the P–A–N framework. (e-i) TEM images of DEX-CeO2. (e-ii) SOD-like activities of DECs. (e-iii) CAT-like activities of DECs. (f) In the repeated UTI model, representative images of bladders and kidneys with/without treatment are shown in (i), and (ii) total bladder and (iii) kidney bacterial counts were determined after homogenization. Reproduced with permission.61 Copyright 2024 American Chemical Society.

Similarly, UTIs represent one of the most prevalent bacterial infections worldwide, frequently caused by uropathogenic Escherichia coli (UPEC) via FimH-mediated adhesion to urinary tract epithelial cells.261 The persistent colonization by UPEC and subsequent inflammatory response result in substantial tissue injury and high recurrence rates. While anti-adhesive glycomimetic agents have been developed to mimic the innate defense mechanism of uromodulin, they often fail to address the accompanying inflammation and oxidative stress, and the excessive use of antibiotics exacerbates inflammation and promotes resistance.262 To overcome these challenges, a bioinspired dextran-modified ceria nanozyme (DEC) has been constructed, offering integrated anti-adhesive and anti-inflammatory properties under the P–A–N framework (Fig. 9(d)). In the preparation phase, ultrasmall ceria nanoparticles are modified with Food and Drug Administration (FDA)-approved dextran, which provides multivalent binding sites mimicking uromodulin glycans for efficient FimH targeting and bacterial capture (Fig. 9(e)-i). In the activation phase, DEC exhibits dual-functionality at the infected site: dextran inhibits UPEC adhesion and colonization, while the ceria core acts as an antioxidant nanozyme, scavenging ROS and suppressing local inflammation triggered by bacterial toxins such as lipopolysaccharides (Fig. 9(e)-i and ii). Finally, in the nontoxic treatment phase, both components of DEC demonstrate high biocompatibility and are safely metabolized, avoiding the drawbacks of antibiotics such as resistance development and secondary inflammatory injury (Fig. 9(f)). This integrated, staged strategy significantly enhances therapeutic efficacy and biosafety in UTI management. Together, these advances in pancatalytic biomaterials employing a preparation–activation–nontoxic treatment paradigm offer a promising and versatile platform for the precise and safe treatment of IDs of the urinary system.61

3.9 Reproductive system

Candida vaginitis, a prevalent gynecological inflammatory disorder caused by Candida albicans infection,263 affects approximately 75% of women globally during their lifetime, with 5–8% developing recurrent infections.264 This condition is characterized by vaginal dysbiosis, local inflammatory responses with epithelial cell damage, and high recurrence rates due to limitations of conventional therapies.265 Current antifungal agents such as clotrimazole and fluconazole demonstrate significant shortcomings including fungistatic rather than fungicidal activity, disruption of normal vaginal microbiota, mucosal damage, and pH imbalance exacerbating recurrence risks.265–267 The FeLab system (rGO@FeS2/Lactobacillus@HA hydrogel) exemplifies an innovative pancatalytic therapeutic strategy based on the P–A–N framework (Fig. 10(a)). In the preparation phase, the system co-encapsulates peroxidase-mimetic rGO@FeS2 nanozymes with probiotic Lactobacillus within a hyaluronic acid hydrogel matrix, enabling pathogen-responsive release through hyaluronidase degradation (Fig. 10(b)-i). During activation, the system simultaneously achieves physiological vaginal pH restoration (4–4.5) through probiotic lactic acid production and targeted pathogen elimination via nanozyme-catalyzed conversion of bacterial-derived H2O2 into hydroxyl radicals (Fig. 10(b)-ii and iii). During the nontoxic treatment phase, the system effectively eradicates C. albicans while preserving beneficial flora, achieves localized therapeutic action that protects mucosal cells, and through probiotic colonization reestablishes a healthy microbiota, thereby significantly reducing recurrence rates (Fig. 10(c)).62 This integrated system demonstrates synergistic catalytic antimicrobial activity and ecological modulation, offering an effective and safe precision therapy that establishes a new paradigm for treating reproductive system inflammatory diseases.
image file: d5nh00408j-f10.tif
Fig. 10 Pancatalytic biomaterials for the treatment of reproductive system diseases. (a) rGO@FeS2 for candida vaginitis via the P–A–N framework. (b-i) TEM images of rGO@FeS2. (b-ii) Ultraviolet-visible absorption spectra and the corresponding color changes of TMB in different reaction systems. (b-iii) EPR monitoring the generation of ˙OH by rGO@FeS2 in the presence of H2O2 in a pH 4.5 HAc–NaAc buffer after 5 min of incubation. (c-i) Digital images of C. albicans colonies in the vaginal washes in different groups. (c-ii) H&E-staining images of the vaginal tissue in different groups. (d-i) Schematic illustration of TPDA NP and PDA NP preparation. Reproduced with permission.62 Copyright 2023, The American Association for the Advancement of Science. (d-ii) TEM image of TPDA NPs. (d-iii) Scanning transmission electron microscopy images with EDS mapping of C, N, and O elemental signals of TPDA NPs. (e-i) SOD-like activity of TPDA NPs. (e-ii) CAT-like ability of TPDA NPs. (e-iii) POD-like ability of TPDA NPs. (e-iv) GPx-like ability of TPDA NPs. (f-i) H&E, Ki-67, and TUNEL staining results of the tumors from different groups. (f-ii) The IHC results of CD36, 3-nitrotyrosine, IL-1β, IL-6, TNF-α, and IL-10 expression in the tumors from the two groups of GO NSs under 808 nm laser irradiation and TPDA NPs under 808 nm laser irradiation. Reproduced with permission.63 Copyright 2024 Wiley-VCH GmbH.

Chronic prostatitis represents a significant global health concern, accounting for over 90% of prostatitis cases with a prevalence of 8.2% worldwide.268 This condition is characterized by persistent inflammatory reactions in the prostate gland, leading to elevated expression of pro-inflammatory cytokines such as IL-1β and TNF-α.269 A key pathological feature is the excessive generation of reactive RO/NSs, including hydroxyl radicals ˙OH, ˙O2, H2O2, NO˙, and ONOO, which disrupt redox homeostasis and induce oxidative stress.270 These pathological changes not only contribute to typical clinical symptoms but also lead to mitochondrial dysfunction and tissue damage. Of particular concern is the established link between chronic prostatitis and more severe complications – long-term oxidative stress and inflammation promote epigenetic alterations that increase the risk of developing prostatic intraepithelial neoplasia and prostate cancer (PCa), the second most common cancer in men globally.271 Additionally, chronic prostatitis significantly impacts male reproductive health, often causing sexual dysfunction and fertility issues.272

Current clinical management of chronic prostatitis encounters multiple challenges, with conventional anti-inflammatory therapies often inducing undesirable systemic side effects and immune modulators potentially provoking unpredictable or excessive immune responses. Moreover, existing treatments fail to adequately address the root cause of oxidative stress or provide comprehensive protection against cancer development. The clinical challenge is further compounded by the need for therapies that can simultaneously manage inflammation, prevent malignant transformation, and preserve reproductive function without causing adverse effects. These unmet medical needs highlight the urgent requirement for innovative treatment strategies that can target multiple pathological pathways in prostate diseases while maintaining an acceptable safety profile. Their therapeutic action follows a P–A–N framework where during preparation, donor–acceptor pairs are constructed to optimize catalytic and photothermal properties (Fig. 10(d)). The activation phase involves catalytic elimination of RO/NSs to restore redox balance coupled with downregulation of CD36 expression to inhibit NF-κB signaling, thereby alleviating inflammation and improving sexual function in experimental autoimmune prostatitis models (Fig. 10(e)). For nontoxic treatment, the system leverages the inherent biocompatibility of polydopamine and 2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO) derivatives while enabling photothermal ablation of cancerous lesions (Fig. 10(f)).63 This integrated approach simultaneously addresses prostatitis symptoms, prevents malignant transformation, and preserves reproductive function, representing a significant advancement over conventional monotherapeutic strategies. The unique combination of catalytic ROS scavenging and photothermal capabilities positions tetramethylpiperidine-1-oxyl doped polydopamine nanoparticles (TPDA NPs) as a promising therapeutic platform for comprehensive prostate disease management.

3.10 Others

Photothermal therapy (PTT) has emerged as a promising tumor treatment modality due to its precise spatiotemporal control and low invasiveness.273 However, the therapeutic efficacy of PTT is significantly compromised by its inherent inflammatory side effects.274 During treatment, localized hyperthermia induces substantial heat stress that triggers a cascade of detrimental biological responses.275 This includes rapid overproduction of ROS, necrotic cell death, and subsequent activation of immune cells. The resulting inflammatory cascade leads to excessive secretion of proinflammatory cytokines (TNF-α, IL-6, IL-1β) and formation of neutrophil extracellular traps, creating an immunosuppressive microenvironment that paradoxically promotes tumor recurrence and metastasis.276 Current PTT approaches face critical limitations in addressing these issues, as they typically focus solely on thermal ablation while neglecting the consequential inflammatory responses, or employ passive anti-inflammatory strategies that only act after inflammatory cascades are already established. To address these challenges, DNA-guided allomelanin nanodisks (DNA-pDHN) were developed as innovative pancatalytic biomaterials. The system operates through a preparation–activation–nontoxic treatment (P–A–N) framework (Fig. 11(a)). In the preparation phase, DHN oligomers are precisely organized into optimized nanostructures (Fig. 11(b)-i). During activation, the nanodisks demonstrate dual photothermal conversion and ROS scavenging capabilities (Fig. 11(b)-ii). This leads to nontoxic treatment outcomes including inflammation control and tumor suppression (Fig. 11(c)). The DNA-pDHN system represents a significant advance in PTT by simultaneously addressing both tumor ablation and inflammatory consequences through integrated catalytic mechanisms.64
image file: d5nh00408j-f11.tif
Fig. 11 Pancatalytic biomaterials for the treatment of other diseases. (a) DNA-pDHN nanodisks for tumor via the P–A–N framework. (b-i) AFM and TEM images of DNA-pDHN. (b-ii) TEM images of DNA-pDHN nanodisks after different treatments. (c-i) Photograph of the tumors collected from different groups of mice after treatment. (c-ii) H&E, Ki67, and TUNEL staining of tumor tissues on day 14. Reproduced with permission.64 Copyright 2023 Wiley-VCH GmbH. (d) UPDA NPs for tumor via the P–A–N framework. (e-i) The step-by-step production and key physical features of UPDA NPs. (e-ii) A representative TEM image of PDA NPs. The inset is a representative photograph of PDA NPs dispersed in deionized water. (e-iii) A representative TEM image of UPDA NPs. (f-i) Weight changes of mice after different treatments. (f-ii) Representative fur appearances of mice receiving different treatments. Reproduced with permission.65 Copyright Published by Elsevier Ltd.

The aging process is marked by a state of persistent, low-grade systemic inflammation, which occurs alongside cellular senescence, decline in immune function (immunosenescence), impaired organ performance, and the onset of age-associated pathologies.277,278 Senescent cells release a complex mixture of signaling molecules, collectively termed the senescence-associated secretory phenotype (SASP), which includes cytokines, chemokines, and growth factors. These SASP components not only sustain chronic inflammatory conditions but also propagate senescence to neighboring healthy cells through paracrine signaling.279 Concurrently, prolonged inflammation accelerates the aging of immune cells, diminishing their functional capacity and impairing their ability to eliminate senescent cells and inflammatory mediators. This establishes a self-perpetuating cycle in which inflammation drives senescence, and senescence, in turn, exacerbates inflammation. Persistently high levels of inflammatory mediators in critical organs, such as the bone marrow, liver, and lungs, contribute to progressive tissue damage and the development of age-related disorders. Given these observations, chronic inflammation is now regarded as an intrinsic contributor to the aging process, and targeting inflammatory pathways may represent a viable therapeutic approach for mitigating age-related decline.280 Recent studies demonstrate that oxidative stress serves as a key molecular link between inflammation and senescence, creating a vicious cycle that accelerates aging. Ultra-small polydopamine nanoparticles (UPDA NPs) represent a paradigm of pancatalytic biomaterials for anti-aging therapy that comprehensively embodies the P–A–N framework (Fig. 11(d)). Engineered through liquid-phase exfoliation, UPDA NPs mimic the cascade activities of natural antioxidant enzymes (CAT/SOD), with their ultrasmall size significantly enhancing radical-scavenging efficiency by maximizing electron-donating capacity and reactive species contact area (Fig. 11(e)). In aging-related microenvironments, these nanoparticles demonstrate broad-spectrum catalytic activation, effectively dismutating ˙O2 to H2O2/O2 (SOD-like activity), decomposing H2O2 into H2O/O2 (CAT-like activity), and directly neutralizing ˙OH, ˙NO, and ONOO, thereby disrupting the oxidative stress–senescence vicious cycle at the molecular level. The system ensures nontoxic treatment through its biocompatible polydopamine matrix and renal-clearable dimensions, while transcriptomic analysis reveals additional anti-senescence effects mediated via NF-κB pathway regulation (Fig. 11(f)). This integrated approach showcases how UPDA NPs synergize catalytic RONS elimination with systemic safety, offering a clinically translatable strategy that holistically addresses aging through synchronized material design (preparation), pathological microenvironment remodeling (activation), and metabolic safety (nontoxic treatment).65

4. Conclusion and perspectives

Pancatalytic biomaterials represent an emerging and transformative paradigm in the treatment of IDs. These engineered materials are defined by their integration of principles from catalytic biology and catalytic medicine, enabling holistic modulation of pathological microenvironments. Within the P–A–N framework, pancatalytic biomaterials function to catalyze specific bio-effects. In inflammatory diseases, their core function lies in either mimicking natural enzymatic activities (e.g., SOD, CAT, GPx) or demonstrating direct ROS-scavenging capabilities, while exhibiting stimuli-responsive catalytic behavior (e.g., to pH, ROS, hypoxia), thereby executing multi-modal bioeffects like oxidative stress regulation, immune cell phenotype reprogramming, inflammation resolution, and tissue homeostasis restoration. The rational design and application of these materials are guided by the P–A–N framework, ensuring targeted action, efficacy, and biocompatibility.

Based on composition and properties, pancatalytic biomaterials are broadly categorized into three main classes: inorganic materials (e.g., CeO2, V2C MXenes, BPNSs), valued for their robust catalytic efficiency and stability; organic materials (e.g., PDA, EGCG), prized for their inherent biocompatibility, structural tunability, and ability to mimic natural enzymes; and organic/inorganic hybrid systems (e.g., metal–organic frameworks, melanin-based composites), which synergistically combine the catalytic potency of inorganic components with the functional versatility and biocompatibility of organic elements. Key characteristics underpinning their therapeutic potential include inherent enzyme-mimetic activities, stimuli-responsiveness for spatiotemporal control, biocompatibility, biodegradability (or controlled metabolic clearance), targeted delivery mechanisms, and inherent multifunctionality allowing theranostic integration. Pancatalytic biomaterials demonstrate broad therapeutic applicability across multiple physiological systems affected by chronic inflammation. They exhibit significant potential in managing inflammatory diseases of the cardiovascular system, nervous system, respiratory system, digestive system, joint system, integumentary system, sensory system, urinary system, reproductive system, and other inflammation-related disorders.

While pancatalytic biomaterials demonstrate substantial therapeutic potential, a critical examination of the P–A–N framework reveals key limitations that must be overcome to realize their full clinical impact. The P–A–N framework, while providing a systematic approach for designing pancatalytic biomaterials, exhibits several inherent limitations that become particularly apparent during clinical translation. The framework's current implementation struggles to adequately account for the dynamic complexity of inflammatory microenvironments, where rapid fluctuations in biochemical markers and substantial spatial heterogeneity often exceed the operational ranges of stimulus-responsive systems, compromising the reliability of the activation phase. Furthermore, the interdependent nature of the preparation, activation, and nontoxic treatment components creates fundamental optimization challenges, where enhancements in catalytic activity frequently come at the expense of biocompatibility, while modifications to improve safety profiles may inadvertently diminish therapeutic efficacy. These limitations are compounded by significant translational gaps between controlled laboratory models and clinical reality, including species-specific metabolic differences and oversimplified disease representations that fail to capture the complexity of human inflammatory conditions. The framework also lacks robust quantitative guidelines for balancing the competing demands of its three components, particularly in multi-catalytic systems designed to address complex, multifactorial inflammatory pathologies. These challenges are most pronounced under chronic inflammatory conditions, where the extended duration of treatment and evolving disease states place exceptional demands on material stability and long-term biocompatibility. These identified limitations, rather than diminishing the framework's value, serve as strategic waypoints guiding the next phase of innovation. The identified limitations of the P–A–N framework, while presenting significant challenges, highlight critical opportunities for advancing pancatalytic biomaterials. This versatile therapeutic platform has demonstrated efficacy across diverse inflammation-related pathologies through multiple breakthroughs, yet remains in its developmental infancy. Systematic solutions must now be developed to overcome these challenges and enable clinical translation (Fig. 12).


image file: d5nh00408j-f12.tif
Fig. 12 Summary of the current applications and the future developments of pancatalytic materials. Created with https://BioRender.com.

4.1. Establishing comprehensive databases

The development of comprehensive databases integrating catalytic performance parameters and disease-specific biomarkers represents a critical frontier in advancing pancatalytic biomaterials for inflammatory diseases. These repositories must systematically characterize catalytic profiles while enabling AI-driven prediction of novel therapeutic pathways. By correlating catalytic signatures with the distinct pathological microenvironments characteristic of chronic inflammation, such as varying pH gradients, ROS profiles, and immune cell distributions, these databases can facilitate personalized biomaterial design tailored to specific inflammatory conditions. The integration of computational modeling with big data analytics allows for intelligent screening of material properties (including size, surface charge, and catalytic kinetics) while predicting their behavior in complex biological systems. This data-driven approach is particularly crucial given the need to balance multi-enzyme catalytic efficiency with fundamental biocompatibility requirements when targeting inflammatory microenvironments. Advanced machine learning algorithms, trained on both material performance data and clinical outcomes, can optimize catalyst design for specific inflammatory pathologies while accounting for critical safety parameters.

Recent successful applications highlight the pivotal role of machine learning and high-throughput screening methodologies: (1) machine learning-assisted identification of RuO2 nanoparticles that combine photothermal tumor ablation with catalase-mimicking anti-inflammatory activity through systematic analysis of structure–activity relationships, demonstrating synergistic therapeutic effects in tumor microenvironments;281 and (2) high-throughput screening-derived SrDy2O4 nanozymes that simultaneously scavenge ROS, modulate gut microbiota through selective bacterial regulation, and provide diagnostic imaging capabilities for ulcerative colitis treatment, showcasing multifunctional therapeutic potential.282 These cases exemplify how machine learning-driven prediction and high-throughput experimental screening in integrated databases synergistically guide the rational design of pancatalytic biomaterials for specific inflammatory conditions. Building on these proof-of-concept demonstrations, such systematic databases promise to transform current empirical approaches into precision platforms for developing next-generation anti-inflammatory therapies. By codifying the relationships between material properties and biological outcomes, these knowledge bases enable the rational design of therapeutics that simultaneously address catalytic potency, microenvironmental responsiveness, and biosafety – ultimately enabling tailored interventions for diverse inflammation-related diseases.

4.2. Elucidating biosafety mechanisms in depth

The clinical translation of pancatalytic biomaterials presents significant biosafety challenges that require comprehensive investigation. While preliminary studies indicate acceptable biocompatibility for certain materials, critical gaps remain in understanding their long-term biological interactions and metabolic fate. Recent advances in material design have introduced innovative strategies to enhance biosafety. For instance, non-metallic catalytic biomaterials like evodiamine, a natural compound with chemotherapeutic and peroxidase-like activity, show promise. When co-delivered with ICG in liposomal nanoparticles, this system enables synergistic chemotherapy/CDT/PDT effects against oral cancer while maintaining excellent biosafety, highlighting the potential of metal-free catalytic systems.283 Similarly, researchers have developed ultrasmall, biodegradable materials such as monodispersed CoS2 nanoclusters. Their loosely stacked structure provides abundant catalytic sites for efficient Fenton-like reactions, converting H2O2 into cytotoxic ·OH in tumors.284 These nanoclusters rapidly dissociate into excretable ions, with toxicological studies confirming minimal toxicity.

The path to clinical translation requires resolving several key challenges. The complex relationship between material properties (surface characteristics, composition, structure) and biological systems raises concerns about inflammatory responses, immune activation, and potential toxicity from degradation byproducts. Current safety assessments are hindered by the lack of standardized evaluation protocols for nanomaterials, particularly regarding their metabolic pathways and tissue-specific accumulation. Thorough preclinical testing must evaluate blood compatibility, multi-organ histopathology, immune responses, biodegradation kinetics, and clearance mechanisms across multiple biological models, including large animals that better approximate human physiology. Key parameters such as optimal dosing, administration routes, and exposure duration need careful optimization to balance therapeutic efficacy with safety. The development of novel catalytic materials with enhanced biocompatibility and controlled degradation profiles, combined with rigorous safety validation through multidisciplinary approaches, will be crucial for advancing these therapies toward clinical applications while ensuring patient safety. This requires establishing comprehensive safety profiles through systematic preclinical studies to gain regulatory approval and build confidence for clinical use.

4.3. Mass production

Mass production of pancatalytic materials faces significant hurdles despite their cost-effective lab-scale synthesis, as complex fabrication procedures and specialized equipment impede scalable manufacturing. Key challenges include the technical difficulty in scaling up ultrasmall or composite nanocatalysts, low stability leading to oxidation during storage, and reliance on toxic solvents that pose contamination risks. Recent breakthroughs in bacterial extracellular vesicle (BEV) production systems offer promising solutions to these challenges. The EXODUS platform (T-2800/H-600) represents a cutting-edge automated system that integrates ultrasound-assisted nanofiltration technology to achieve high-throughput, GMP-compliant isolation of BEVs (20–400 nm) with remarkable purity and batch-to-batch consistency. Notably, these naturally derived BEVs demonstrate potent ROS scavenging capabilities, making them promising candidates for anti-inflammatory applications.285,286 Their unique lipid bilayer structure enables efficient loading of therapeutic cargo. Similarly, the development of EGCG polymer colloidal spheres through an innovative combination of ternary condensation reactions and ethanol-mediated nanoprecipitation has demonstrated exceptional drug loading capacities (up to 86% for paclitaxel) while preserving catalytic activity. The integration of continuous microfluidic processing (5 mL min−1) in this system ensures precise control over particle characteristics (PDI < 0.2, size < 200 nm), addressing critical challenges in nanoparticle manufacturing consistency.287

To enable the clinical translation of pancatalytic materials, standardized protocols must be developed that incorporate lessons from these successful examples, focusing on reproducible synthesis under mild, green conditions while eliminating organic solvent residues and ensuring rigorous quality control of physical properties (size, morphology, surface chemistry). Production processes must be further optimized through collaborative efforts across academia, industry, and regulatory bodies to establish comprehensive manufacturing guidelines that address cost reduction, workflow simplification, and stringent quality control requirements that can effectively bridge the gap between bench-scale innovation and clinical deployment.

4.4. Clinical translation

The clinical translation of pancatalytic biomaterials requires establishing robust evaluation systems and comprehensive clinical validation. Current preclinical models often show limited predictive value for human therapeutic outcomes, highlighting the need for more physiologically relevant testing platforms across multiple biological systems. Thorough clinical investigations must systematically assess safety profiles, treatment efficacy, and long-term performance to build a solid evidence base for regulatory approval. Successful implementation will depend on collaborative efforts between research institutions, clinical centers, and industry partners to bridge the gap between laboratory development and practical applications. Parallel educational initiatives should enhance the understanding of these novel therapies among medical professionals and patients. By addressing these fundamental aspects through coordinated research and development strategies, pancatalytic biomaterials can progress toward becoming clinically viable solutions for managing complex inflammatory conditions.

Building upon these critical considerations, sustained multidisciplinary efforts must be directed toward achieving fundamental breakthroughs in pancatalytic biomaterials to establish this innovative approach as a frontline therapeutic strategy for inflammatory diseases. While significant challenges remain in database establishment, biosafety elucidation, mass production, and clinical translation, these hurdles precisely define the roadmap for future research and development. Through systematic optimization of the P–A–N framework and standardization of interdisciplinary evaluation protocols, pancatalytic biomaterials are poised to revolutionize inflammatory disease management by synergizing targeted catalytic efficacy with enhanced biocompatibility. We anticipate that continued innovation in bioinspired material design, coupled with advanced manufacturing technologies and rigorous clinical validation, will accelerate the transition of these smart therapeutic systems from laboratory breakthroughs to clinical reality. The convergence of materials science, nanotechnology, and immunology embodied by pancatalytic biomaterials promises to open new frontiers in precision medicine, ultimately delivering safer and more effective treatments for patients suffering from inflammatory conditions worldwide.

Author contributions

Xiaoyan Jiang: investigation, data curation, visualization, writing – original draft. Yu Chen: conceptualization, funding acquisition, supervision, writing – review & editing.

Conflicts of interest

There are no conflicts to declare.

Abbreviations

IDInflammatory disease
PPreparation
AActivation
NNontoxic treatment
IDsInflammation-related diseases
P–A–NPreparation of the catalyst, activation of the biological effect and nontoxic treatment of diseases
IBDInflammatory bowel disease
ASAtherosclerosis
MSMultiple sclerosis
SNHLSensorineural hearing loss
PAMPsPathogen-associated molecular patterns
DAMPsDamage-associated molecular patterns
ROSReactive oxygen species
NSAIDsNonsteroidal anti-inflammatory drugs
GaHCF NACsGaHCF nanoabsorption catalysts
MIMyocardial infarction
RARheumatoid arthritis
CSNsCalcium disilicide nanoparticles
SODSuperoxide dismutase
CATCatalase
GPxGlutathione peroxidase
CaH NCsHexacyanoferrate nanocatalysts
CeO2Cerium oxide
Ce NPsCeria nanoparticles
MSNsMesoporous silica nanoparticles
BPNSsBlack phosphorus nanosheets
PDAPolydopamine
PICsomesPolyion complex vesicles
EGCGEpigallocatechin gallate
MCNMelanin-based composites
CVDsCardiovascular diseases
WHOWorld Health Organization
LDLLow-density lipoprotein
ox-LDLOxidized LDL
IL-6Interleukin-6
TNF-αTumor necrosis factor-α
RONSReactive oxygen/nitrogen species
MIMMyocardial infarction microenvironment
Pt NCsPlatinum nanoclusters
BBBBlood–brain barrier
Apo-LFApo-lactoferrin
˙O2Superoxide anion
H2O2Hydrogen peroxide
˙OHHydroxyl radicals
NONitric oxide
ONOOPeroxynitrite
NADPHNicotinamide adenine dinucleotide phosphate
NOXNADPH oxidase
CaHF NPsCalcium hexacyanoferrate(III) nanoparticles
CNSCentral nervous system
ADAlzheimer's disease
PDParkinson's disease
HDHuntington's disease
MSMultiple sclerosis
IL-1βInterleukin-1β
NLRP3NOD-like receptor family pyrin domain containing 3
NF-κBNuclear factor kappa-light-chain-enhancer of activated B cells
IL-10Interleukin-10
JAK-STATJanus kinase-signal transducer and activator of transcription
MAPKMitogen-activated protein kinase
TLRsToll-like receptors
TGF-βTransforming growth factor-beta
β-Amyloid
GCAGuanidinium-modified calixarene
APAscorbyl palmitate
IP6Dipotassium phytate
DADopamine
α-synα-Synuclein
CMAChaperone-mediated autophagy
DBSDeep brain stimulation
TRPV1Transient receptor potential vanilloid family member 1
SNSubstantia nigra
β-synβ-Synuclein
polyQPolyglutamine
HttHuntingtin
mHttMutant Htt
NPsNanoparticles
mHtt-Q7474 polyglutamine repeats
UPSUbiquitin-proteasome system
APCsAntigen-presenting cells
COPDChronic obstructive pulmonary disease
CXCL8C–X–C motif ligand 8
CCL2C–C motif ligand 2
PEGPolyethylene glycol
BPQDsBlack phosphorus quantum dots
CAT-NGsCAT-loaded nanogels
MASHMetabolic dysfunction-associated steatohepatitis
IBDsInflammatory bowel diseases
CDCrohn's disease
UCUlcerative colitis
LNPsLipid nanoparticles
RBP-4Retinol-binding protein-4
HSCsHepatic stellate cells
CBDCollagen-binding domain
ECMExopolysaccharide-rich extracellular matrix
OAOsteoarthritis
RARheumatoid arthritis
DMARDsDisease-modifying anti-rheumatic drugs
CAPsCalcium phosphate particles
H2Hydrogen gas
ADAtopic dermatitis
HCHair cell
NAC CPDsN-Acetyl-L-cysteine-derived carbonized polymer dots
BAXBCL2-associated X protein
DEDDry eye disease
DSPDexamethasone sodium phosphate
AKIAcute kidney injury
UTIsUrinary tract infections
UPECUropathogenic Escherichia coli
DECDextran-modified ceria nanozyme
FDAFood and Drug Administration
PCaProstate cancer
TEMPO2,2,6,6-Tetramethylpiperidine-1-oxyl
TPDA NPsTetramethylpiperidine-1-oxyl doped polydopamine nanoparticles
PTTPhotothermal therapy
DNA-pDHNDNA-guided allomelanin nanodisks
CurCurcumin
TATannic acid
DSSDextran sulfate sodium
FADFamilial Alzheimer's disease
ACLTAnterior cruciate ligament transection
SASPSenescence-associated secretory phenotype
BAKBenzalkonium chloride
UPDA NPsUltra-small polydopamine nanoparticles
DCFHDSP@Ce-MOFs F127 hydrogel
MLPGaLyticase and gallium ions co-integrated nanosystem

Data availability

No primary research results, software or code have been included and no new data were generated or analysed as part of this review.

Acknowledgements

We thank the funding support of the National Key Research and Development Projects (grants no. 2023YFC2306500), Shanghai Science and Technology Program (grant no. 24520712900), and “Shuguang Program” supported by the Shanghai Education Development Foundation and Shanghai Municipal Education Commission (grant no. 21SG39).

References

  1. R. Medzhitov, Nature, 2008, 454, 428–435 CrossRef CAS PubMed.
  2. P. J. Barnes, F. H. Epstein and M. Karin, N. Engl. J. Med., 1997, 336, 1066–1071 CrossRef CAS PubMed.
  3. G. K. Hansson, N. Engl. J. Med., 2005, 353, 429–430 CrossRef CAS PubMed.
  4. G. Schett and M. F. Neurath, Nat. Commun., 2018, 9, 3261 CrossRef PubMed.
  5. D. Furman, J. Campisi, E. Verdin, P. Carrera-Bastos, S. Targ, C. Franceschi, L. Ferrucci, D. W. Gilroy, A. Fasano, G. W. Miller, A. H. Miller, A. Mantovani, C. M. Weyand, N. Barzilai, J. J. Goronzy, T. A. Rando, R. B. Effros, A. Lucia, N. Kleinstreuer and G. M. Slavich, Nat. Med., 2019, 25, 1822–1832 Search PubMed.
  6. J. Zindel and P. Kubes, Annu. Rev. Pathol.: Mech. Dis., 2020, 15, 493–518 CrossRef CAS PubMed.
  7. A. D. Luster, R. Alon and U. H. von Andrian, Nat. Immunol., 2005, 6, 1182–1190 CrossRef CAS PubMed.
  8. P. Rajendran, Y. F. Chen, Y. F. Chen, L. C. Chung, S. Tamilselvi, C. Y. Shen, C. H. Day, R. J. Chen, V. P. Viswanadha, W. W. Kuo and C. Y. Huang, J. Cell. Physiol., 2018, 233, 6458–6471 CrossRef CAS PubMed.
  9. Z. X. Tu, Y. L. Zhong, H. Z. Hu, D. Shao, R. N. Haag, M. Schirner, J. Lee, B. Sullenger and K. W. Leong, Nat. Rev. Mater., 2022, 7, 557–574 CrossRef CAS PubMed.
  10. K. L. Rock, E. Latz, F. Ontiveros and H. Kono, Annu. Rev. Immunol., 2010, 28, 321–342 CrossRef CAS PubMed.
  11. M. Antonelli and I. Kushner, FASEB J., 2017, 31, 1787–1791 CrossRef CAS PubMed.
  12. R. Chovatiya and R. Medzhitov, Mol. Cell, 2014, 54, 281–288 CrossRef CAS PubMed.
  13. E. A. Azzopardi, E. L. Ferguson and D. W. Thomas, J. Antimicrob. Chemother., 2013, 68, 257–274 CrossRef CAS PubMed.
  14. F. Yuan, L. D. Quan, L. Cui, S. R. Goldring and D. Wang, Adv. Drug Delivery Rev., 2012, 64, 1205–1219 CrossRef CAS PubMed.
  15. I. Tabas and C. K. Glass, Science, 2013, 339, 166–172 CrossRef CAS PubMed.
  16. H. Huang and Y. Chen, Sci. Bull., 2025, 70, 1554–1558 CrossRef CAS PubMed.
  17. X. Song, L. Yu, L. Chen and Y. Chen, Acc. Mater. Res., 2024, 5, 271–285 CrossRef CAS.
  18. H. J. Wang, W. J. He, J. Liao, S. S. Wang, X. Y. Dai, M. H. Yu, Y. J. Xie and Y. Chen, Adv. Mater., 2025, 37, 2411967 CrossRef CAS PubMed.
  19. X. Y. Jiang, M. Y. Yang, Y. Fang, Z. Y. Yang, X. Y. Dai, P. Gu, W. Feng and Y. Chen, Adv. Healthcare Mater., 2023, 12, 2300699 CrossRef CAS PubMed.
  20. M. Y. Yang, X. Y. Jiang, M. Y. Song, W. Feng, Y. Chen, P. Gu and X. Q. Fan, Adv. Funct. Mater., 2025, 35, 2415688 Search PubMed.
  21. W. Feng, X. G. Han, H. Hu, M. Q. Chang, L. Ding, H. J. Xiang, Y. Chen and Y. H. Li, Nat. Commun., 2021, 12, 2203 CrossRef CAS PubMed.
  22. Y. M. Liu, S. Y. Wang, J. X. Zhang, Q. Sun, Y. Xiao, J. Chen, M. L. Yao, G. G. Zhang, Q. Huang, T. J. Zhao, Q. Huang, X. J. Shi, C. Feng, K. L. Ai and Y. P. Bai, Nat. Commun., 2024, 15, 6651 CrossRef CAS PubMed.
  23. S. Wang, X. Liu, Y. Wu, H. Huang, X. Song, W. Feng, X. Dai and Y. Chen, Cell Biomater., 2025, 100101,  DOI:10.1016/j.celbio.2025.100101.
  24. W. Q. Pan, S. H. Shao, X. Y. Cao, X. Y. Dai, Y. Zheng, J. Y. Cheng, W. Feng, R. Wu and Y. Chen, J. Am. Chem. Soc., 2025, 147, 3553–3569 CrossRef CAS PubMed.
  25. P. H. Ji, S. W. Qiu, J. C. Huang, L. P. Wang, Y. M. Wang, P. Wu, M. F. Huo and J. L. Shi, Adv. Mater., 2025, 37, e2415543 CrossRef PubMed.
  26. S. N. Feng, Y. Gong, L. L. Xia, Y. Lang, Y. Z. Shen, H. Li, W. Feng, F. X. Chen and Y. Chen, Adv. Mater., 2024, 36, e2405655 CrossRef PubMed.
  27. Q. Z. Xuan, J. Z. Cai, Y. Gao, X. C. Qiao, T. H. Jin, M. Peydayesh, J. T. Zhou, Q. Y. Sun, L. J. Zhan, B. Liu, P. Wang, H. Li, C. Chen and R. Mezzenga, Adv. Mater., 2025, 37, e2417774 CrossRef PubMed.
  28. X. Hu, F. Y. Li, F. Xia, X. Guo, N. Wang, L. L. Liang, B. Yang, K. L. Fan, X. Y. Yan and D. S. Ling, J. Am. Chem. Soc., 2020, 142, 1636–1644 CrossRef CAS PubMed.
  29. X. Z. Zhang, S. C. Lin, R. Huang, A. Gupta, S. Fedeli, R. Cao-Milan, D. C. Luther, Y. C. Liu, M. D. Jiang, G. T. Li, B. Rondon, H. Wei and V. M. Rotello, J. Am. Chem. Soc., 2022, 144, 12893–12900 CrossRef CAS PubMed.
  30. L. Y. Fu, Y. Zhang, R. A. Farokhzad, B. B. Mendes, J. Conde and J. J. Shi, Chem. Soc. Rev., 2023, 52, 7579–7601 RSC.
  31. H. N. Wang, Z. J. Xu, M. Zhao, G. T. Liu and J. Wu, Biomater. Sci., 2021, 9, 1530–1546 RSC.
  32. Z. J. Xu, S. Y. Han, Z. P. Gu and J. Wu, Adv. Healthcare Mater., 2020, 9, 1901502 CrossRef CAS PubMed.
  33. D. K. Min, Y. E. Kim, M. K. Kim, S. W. Choi, N. Park and J. Kim, ACS Nano, 2023, 17, 24404–24416 CrossRef CAS PubMed.
  34. X. L. Liu, B. H. Chen, J. Q. Chen, X. Wang, X. F. Dai, Y. Q. Li, H. Y. Zhou, L. M. Wu, Z. Liu and Y. Yang, Adv. Mater., 2024, 36, e2308477 CrossRef PubMed.
  35. H. B. Wu, F. Xia, L. X. Zhang, C. Y. Fang, J. Y. Lee, L. J. Gong, J. Q. Gao, D. S. Ling and F. Y. Li, Adv. Mater., 2022, 34, e2108348 CrossRef PubMed.
  36. W. J. Le, Z. Y. Sun, T. Y. Li, H. Cao, C. X. Yang, T. X. Mei, L. H. Zhang, Y. B. Wang, W. W. Jia, W. Sun, Y. H. Hu and Z. M. Liu, Adv. Funct. Mater., 2024, 34, 2314328 CrossRef CAS.
  37. H. L. Lu, J. H. Wei, K. Y. Liu, Z. H. Li, T. Y. Xu, D. Yang, Q. M. Gao, H. J. Xiang, G. D. Li and Y. Chen, ACS Nano, 2023, 17, 6131–6146 CrossRef CAS PubMed.
  38. Y. Wu, W. Q. Xu, L. Jiao, W. L. Gu, D. Du, L. Y. Hu, Y. H. Lin and C. Z. Zhu, Chem. Soc. Rev., 2022, 51, 6948–6964 RSC.
  39. P. Y. Wen, A. Dirisala, H. C. Guo, X. Y. Liu, S. Kobayashi, H. Kinoh, T. Anada, M. Tanaka, K. Kataoka and J. J. Li, J. Controlled Release, 2025, 382, 113683 CrossRef CAS PubMed.
  40. X. Zuo, X. P. Guo, Y. N. Gu, D. Zhao, Z. Zou, Y. Y. Shen, C. L. He, Y. Rong, C. N. Xu and F. Wang, Adv. Funct. Mater., 2024, 34, 2316496 CrossRef CAS.
  41. S. Y. Xian, F. C. Meng, X. N. Chen, L. Q. Zhu and H. X. Wang, J. Nanobiotechnol., 2025, 23, 247 CrossRef CAS PubMed.
  42. Y. Y. Zheng, X. J. Chen, Q. Zhang, L. Yang, Q. Chen, Z. Chen, Y. Wang and D. Wu, Adv. Healthcare Mater., 2024, 13, e2401816 CrossRef PubMed.
  43. X. T. Lou, Y. Y. Hu, H. Zhang, J. Liu and Y. Zhao, J. Nanobiotechnol., 2021, 19, 436 CrossRef CAS PubMed.
  44. Y. W. Hu, Y. H. Miao, Y. Zhang, X. Wang, X. Liu, W. Zhang and D. W. Deng, ACS Nano, 2024, 18, 27557–27569 CrossRef CAS PubMed.
  45. W. Y. Cui, C. Sheng, T. Y. Hu, T. L. Zhou, Q. Chen, L. Y. Jiang, X. Y. Cheng, J. Ji, K. Yao and H. J. Han, ACS Nano, 2024, 18, 11084–11102 CrossRef CAS PubMed.
  46. R. Wang, W. M. Nie, X. Yan, K. K. Luo, Q. Zhang, T. Wang, E. H. Lu, Y. T. Chen, Y. Luo, Z. W. Zhang, H. Wang, J. Zhao and X. Y. Sha, Adv. Mater., 2025, 37, e2409176 CrossRef PubMed.
  47. X. X. Ye, L. L. Xia, H. Yang, J. Y. Xu, T. F. Liu, L. P. Wang, S. Zhang, Y. Chen, D. S. Du and W. Feng, Mater. Today, 2023, 68, 148–163 CrossRef CAS.
  48. W. B. Li, L. L. Xu, S. L. Wang, Y. Y. Wang, Y. C. Pan, L. Q. Shi and D. S. Guo, Adv. Mater., 2024, 36, e2401918 CrossRef PubMed.
  49. J. G. Wu, X. J. Cui, L. Bao, G. Y. Liu, X. Y. Wang and C. Y. Chen, Sci. Adv., 2025, 11, eado4927 CrossRef CAS PubMed.
  50. L. Zhang, P. F. Wei, Y. H. Song, L. Dong, Y. D. Wu, Z. Y. Hao, S. Fan, S. Tai, J. L. Meng, Y. Lu, J. Z. Xue, C. Z. Liang and L. P. Wen, Biomaterials, 2019, 216, 119248 CrossRef CAS PubMed.
  51. T. L. Nguyen, Y. Choi, J. Im, H. Shin, N. M. Phan, M. K. Kim, S. W. Choi and J. Kim, Nat. Commun., 2022, 13, 7449 CrossRef CAS PubMed.
  52. Z. B. Li, G. H. Luo, W. P. Hu, J. L. Hua, S. Y. Geng, P. K. Chu, J. Zhang, H. Y. Wang and X. F. Yu, Angew. Chem., Int. Ed., 2020, 59, 20568–20576 CrossRef CAS PubMed.
  53. Z. X. Chen, Y. X. Zhu, Y. L. You, M. Ge, Y. H. Chen, H. Lin and J. L. Shi, J. Am. Chem. Soc., 2025, 147, 13424–13436 CrossRef CAS PubMed.
  54. X. Z. Shan, Z. Q. Zhao, P. P. Lai, Y. X. Liu, B. Y. Li, Y. B. Ke, H. Q. Jiang, Y. L. Zhou, W. Z. Li, Q. Wang, P. X. Qin, Y. Z. Xue, Z. H. Zhang, C. L. Wei, B. Ma, W. Liu, C. Luo, X. G. Lu, J. Q. Lin, L. Shu, Y. Jie, X. D. Xian, D. Delcassian, Y. F. Ge and L. Miao, Nat. Commun., 2024, 15, 7263 CrossRef CAS PubMed.
  55. F. Huang, X. Y. Lu, L. Kuai, Y. Ru, J. S. Jiang, J. K. Song, S. Chen, L. J. Mao, Y. Y. Li, B. Li, H. Q. Dong and J. L. Shi, J. Am. Chem. Soc., 2024, 146, 3186–3199 CrossRef CAS PubMed.
  56. X. Y. Lu, L. Kuai, F. Huang, J. S. Jiang, J. K. Song, Y. Q. Liu, S. Chen, L. J. Mao, W. Peng, Y. Luo, Y. Y. Li, H. Q. Dong, B. Li and J. L. Shi, Nat. Commun., 2023, 14, 6767 CrossRef CAS PubMed.
  57. S. J. Ge, A. D. Sun, X. Y. Zhou, P. Niu, Y. Chen, X. T. Bao, M. Yu, Z. H. Zhong, J. W. Sun and G. Li, Adv. Mater., 2025, 37, e2413371 CrossRef PubMed.
  58. X. Y. Liu, Z. X. Chen, J. Y. Bai, X. Li, X. Chen, Z. Li, H. X. Pan, S. H. Li, Q. Y. Gao, N. Zhao, A. D. Chen, H. L. Xu, Y. N. Wen, L. Du, M. Yang, X. T. Zhou and J. H. Huang, ACS Nano, 2023, 17, 25377–25390 CrossRef CAS PubMed.
  59. J. He, Y. Ye, D. X. Zhang, K. Yao and M. Zhou, Adv. Mater., 2022, 34, e2206437 CrossRef PubMed.
  60. W. Jiang, X. Hou, Y. Qi, Z. Wang, Y. Liu, X. J. Gao, T. Wu, J. Guo, K. Fan and W. Shang, Adv. Sci., 2024, 11, 2303901 CrossRef CAS PubMed.
  61. Y. H. Zhang, W. L. Liu, G. Wei, Q. Y. Liu, G. Q. Shao, X. Gu, X. M. Cui, Z. J. Zhou, Y. T. Wang, S. Zhao, F. Muhammad, S. R. Li, T. Li, Y. Du and H. Wei, ACS Nano, 2024, 18, 9019–9030 CrossRef CAS PubMed.
  62. G. Wei, Q. Y. Liu, X. Y. Wang, Z. J. Zhou, X. Z. Zhao, W. Q. Zhou, W. L. Liu, Y. H. Zhang, S. J. Liu, C. X. Zhu and H. Wei, Sci. Adv., 2023, 9, adg0949 CrossRef PubMed.
  63. K. Y. Wang, W. P. Mao, X. R. Song, W. Feng, G. C. Wang, T. Zhang, T. Y. Xu, M. Chen, Y. Chen and B. Peng, Adv. Funct. Mater., 2024, 34, 2313528 CrossRef CAS.
  64. S. Wang, J. X. Huang, H. Y. Zhu, J. Zhu, Z. Q. Wang, Y. X. Xing, X. Y. Xie, K. Y. Cai and J. X. Zhang, Adv. Funct. Mater., 2023, 33, 2213151 CrossRef CAS.
  65. J. M. Han, J. Wang, H. W. Shi, Q. Li, S. B. Zhang, H. Wu, W. J. Li, L. H. Gan, H. M. Brown-Borg, W. Feng, Y. Chen and R. C. Zhao, Mater. Today Bio, 2023, 19, 100544 CrossRef CAS PubMed.
  66. W. C. Poller, M. Nahrendorf and F. K. Swirski, Circ. Res., 2020, 126, 1061–1085 CrossRef CAS PubMed.
  67. J. Sanz and Z. A. Fayad, Nature, 2008, 451, 953–957 CrossRef CAS PubMed.
  68. G. A. Roth, G. A. Mensah, C. O. Johnson, G. Addolorato, E. Ammirati, L. M. Baddour, N. C. Barengo, A. Z. Beaton, E. J. Benjamin, C. P. Benziger, A. Bonny, M. Brauer, M. Brodmann, T. J. Cahill, J. Carapetis, A. L. Catapano, S. S. Chugh, L. T. Cooper, J. Coresh, M. Criqui, N. DeCleene, K. A. Eagle, S. Emmons-Bell, V. L. Feigin, J. Fernández-Solà, G. Fowkes, E. Gakidou, S. M. Grundy, F. J. He, G. Howard, F. Hu, L. Inker, G. Karthikeyan, N. Kassebaum, W. Koroshetz, C. Lavie, D. Lloyd-Jones, H. S. Lu, A. Mirijello, A. M. Temesgen, A. Mokdad, A. E. Moran, P. Muntner, J. Narula, B. Neal, M. Ntsekhe, G. M. de Oliveira, C. Otto, M. Owolabi, M. Pratt, S. Rajagopalan, M. Reitsma, A. L. P. Ribeiro, N. Rigotti, A. Rodgers, C. Sable, S. Shakil, K. Sliwa-Hahnle, B. Stark, J. Sundström, P. Timpel, I. M. Tleyjeh, M. Valgimigli, T. Vos, P. K. Whelton, M. Yacoub, L. Zuhlke, C. Murray, V. Fuster and G. N. Jacc, J. Am. Coll. Cardiol., 2020, 76, 2982–3021 CrossRef PubMed.
  69. E. J. Benjamin, P. Muntner, A. Alonso, M. S. Bittencourt, C. W. Callaway, A. P. Carson, A. M. Chamberlain, A. R. Chang, S. Cheng, S. R. Das, F. N. Delling, L. Djousse, M. S. V. Elkind, J. F. Ferguson, M. Fornage, L. C. Jordan, S. S. Khan, B. M. Kissela, K. L. Knutson, T. W. Kwan, D. T. Lackland, T. T. Lewis, J. H. Lichtman, C. T. Longenecker, M. S. Loop, P. L. Lutsey, S. S. Martin, K. Matsushita, A. E. Moran, M. E. Mussolino, M. O’Flaherty, A. Pandey, A. M. Perak, W. D. Rosamond, G. A. Roth, U. K. A. Sampson, G. M. Satou, E. B. Schroeder, S. H. Shah, N. L. Spartano, A. Stokes, D. L. Tirschwell, C. W. Tsao, M. P. Turakhia, L. B. VanWagner, J. T. Wilkins, S. S. Wong, S. S. Virani, A. H. A. C. Epid, P. S. Comm and S. S. Subcomm, Circulation, 2019, 139, E56–E528 CrossRef PubMed.
  70. S. Barquera, A. Pedroza-Tobías, C. Medina, L. Hernández-Barrera, K. Bibbins-Domingo, R. Lozano and A. E. Moran, Arch. Med. Res., 2015, 46, 328–338 CrossRef PubMed.
  71. B. A. Kaufmann, J. M. Sanders, C. Davis, A. Xie, P. Aldred, I. J. Sarembock and J. R. Lindner, Circulation, 2007, 116, 276–284 CrossRef CAS PubMed.
  72. D. P. Ramji and T. S. Davies, Cytokine Growth Factor Rev., 2015, 26, 673–685 CrossRef CAS PubMed.
  73. M. Kim, A. Sahu, Y. Hwang, G. B. Kim, G. H. Nam, I. S. Kim, I. C. Kwon and G. Tae, Biomaterials, 2020, 226, 119550 CrossRef CAS PubMed.
  74. C. Weber and H. Noels, Nat. Med., 2011, 17, 1410–1422 CrossRef CAS PubMed.
  75. N. MacRitchie, V. Di Francesco, M. F. M. M. Ferreira, T. J. Guzik, P. Decuzzi and P. Maffia, Semin. Immunol., 2021, 56, 101536 CrossRef CAS PubMed.
  76. V. Garlapati, M. Molitor, T. Michna, G. S. Harms, S. Finger, R. Jung, J. Lagrange, P. Efentakis, J. Wild, M. Knorr, S. Karbach, S. Wild, K. Vujacic-Mirski, T. Muenzel, A. Daiber, M. Brandt, T. Gori, H. Milting, S. Tenzer, W. Ruf and P. Wenzel, J. Clin. Invest., 2023, 133, e156436 CrossRef CAS PubMed.
  77. T. J. Beldman, T. S. Malinova, E. Desclos, A. E. Grootemaat, A. L. S. Misiak, S. van der Velden, C. P. A. A. van Roomen, L. Beckers, H. A. van Veen, P. M. Krawczyk, R. A. Hoebe, J. C. Sluimer, A. E. Neele, M. P. J. de Winther, N. N. van der Wel, E. Lutgens, W. J. M. Mulder, S. Huveneers and E. Kluza, ACS Nano, 2019, 13, 13759–13774 CrossRef CAS PubMed.
  78. J. Liu and J. Wu, Am. J. Hypertens., 2024, 37, 241–247 CrossRef CAS PubMed.
  79. D. P. Leong, P. G. Joseph, M. McKee, S. S. Anand, K. K. Teo, J. D. Schwalm and S. Yusuf, Circ. Res., 2017, 121, 695–710 CrossRef CAS PubMed.
  80. D. Zhao, J. Liu, M. Wang, X. G. Zhang and M. G. Zhou, Nat. Rev. Cardiol., 2019, 16, 203–212 CrossRef PubMed.
  81. P. Libby, Nature, 2021, 592, 524–533 CrossRef CAS PubMed.
  82. Z. A. Fayad, P. M. Robson and V. Fuster, J. Am. Coll. Cardiol., 2024, 83, 2145–2147 CrossRef PubMed.
  83. A. Arbab-Zadeh and V. Fuster, J. Am. Coll. Cardiol., 2015, 65, 846–855 CrossRef PubMed.
  84. W. J. Zhang, H. Wei and B. Frei, Exp. Biol. Med., 2010, 235, 633–641 CrossRef CAS PubMed.
  85. F. Vinchi, G. Porto, A. Simmelbauer, S. Altamura, S. T. Passos, M. Garbowski, A. M. N. Silva, S. Spaich, S. E. Seide, R. Sparla, M. W. Hentze and M. U. Muckenthaler, Eur. Heart J., 2020, 41, 2681–2695 CrossRef CAS PubMed.
  86. P. M. Ridker, B. M. Everett, T. Thuren, J. G. MacFadyen, W. H. Chang, C. Ballantyne, F. Fonseca, J. Nicolau, W. Koenig, S. D. Anker, J. J. P. Kastelein, J. H. Cornel, P. Pais, D. Pella, J. Genest, R. Cifkova, A. Lorenzatti, T. Forster, Z. Kobalava, L. Vida-Simiti, M. Flather, H. Shimokawa, H. Ogawa, M. Dellborg, P. R. F. Rossi, R. P. T. Troquay, P. Libby, R. J. Glynn and C. T. Grp, N. Engl. J. Med., 2017, 377, 1119–1131 CrossRef CAS PubMed.
  87. S. M. Nidorf, E. Ben-Chetrit and P. M. Ridker, Eur. Heart J., 2024, 45, 1596–1601 CrossRef CAS PubMed.
  88. S. Khan, H. Bilal, M. N. Khan, W. J. Fang, W. Q. Chang, B. Yin, N. J. Song, Z. R. Liu, D. X. Zhang, F. Yao, X. Wang, Q. Wang, L. Cai, B. Hou, J. Y. Wang, C. Y. Mao, L. X. Liu and Y. B. Zeng, Front. Immunol., 2024, 15, 1372693 CrossRef CAS PubMed.
  89. D. Weissman and C. Maack, J. Clin. Invest., 2023, 133, e167079 CrossRef CAS PubMed.
  90. A. D. Wilson, M. A. Richards, M. K. Curtis, M. Rohling, S. Monterisi, A. A. Loonat, J. Miller, V. Ball, A. Lewis, D. Tyler, A. Moshnikova, O. A. Andreev, Y. K. Reshetnyak, C. Carr and P. Swietach, Cardiovasc. Res., 2022, 118, 2946–2959 CrossRef CAS PubMed.
  91. M. A. Redd, S. E. Scheuer, N. J. Saez, Y. Yoshikawa, H. S. Chiu, L. Gao, M. Hicks, J. E. Villanueva, Y. Joshi, C. Y. Chow, G. Cuellar-Partida, J. N. Peart, L. E. S. Hoe, X. L. Chen, Y. L. Z. Sun, J. Y. Suen, R. J. Hatch, B. Rollo, D. Xia, M. A. H. Alzubaidi, S. Maljevic, G. A. Quaife-Ryan, J. E. Hudson, E. R. Porrello, M. Y. White, S. J. Cordwell, J. F. Fraser, S. Petrou, M. E. Reichelt, W. G. Thomas, G. F. King, P. S. Macdonald and N. J. Palpant, Circulation, 2021, 144, e211 CrossRef PubMed.
  92. X. Y. Long, M. Liu, Y. Y. Nan, Q. H. Chen, Z. X. Xiao, Y. T. Xiang, X. H. Ying, J. Sun, Q. Huang and K. L. Ai, Adv. Mater., 2024, 36, e2308239 CrossRef PubMed.
  93. Y. Guan, H. Niu, J. X. Wen, Y. Dang, M. Zayed and J. J. Guan, ACS Nano, 2022, 16, 19551–19566 CrossRef CAS PubMed.
  94. Q. Sun, H. Q. Ma, J. X. Zhang, B. Y. You, X. H. Gong, X. L. Zhou, J. Chen, G. G. Zhang, J. Huang, Q. Huang, Y. R. Yang, K. L. Ai and Y. P. Bai, Adv. Sci., 2023, 10, e2204999 CrossRef PubMed.
  95. M. Shilo, H. Oved, L. Wertheim, I. Gal, N. Noor, O. Green, E. S. Baruch, D. Shabat, A. Shapira and T. Dvir, Adv. Sci., 2021, 8, e2102919 CrossRef PubMed.
  96. Q. Chen, J. Wang, X. X. Xiong, J. Y. Chen, B. Wang, H. X. Yang, J. L. Zhou, H. P. Deng, L. J. Gu and J. Tian, Adv. Healthcare Mater., 2024, e2402376 CrossRef PubMed.
  97. R. Q. Cheng, X. Q. Luo, X. H. Wu, Z. J. Wang, Z. Q. Chen, S. R. Zhang, H. M. Xiao, J. J. Zhong, R. R. Zhang, Y. Cao and X. Y. Qin, Adv. Sci., 2024, 11, e2410529 CrossRef PubMed.
  98. GBD 2019 Stroke Collaborators, Lancet Neurol., 2021, 20, 795–820 CrossRef PubMed.
  99. M. L. Jia, F. Y. Jin, S. J. Li, C. H. Ren, M. Ruchi, Y. C. Ding, W. B. Zhao and X. M. Ji, CNS Neurosci. Ther., 2024, 30, e14631 CrossRef PubMed.
  100. X. H. Zhang, X. Y. Wang, Y. Q. Li, Y. Y. Zhang, H. Zhu, C. Xie, Y. D. Zhou, Y. Shen and J. P. Tong, Cell. Mol. Neurobiol., 2024, 44, 19 CrossRef CAS PubMed.
  101. H. W. Zheng, X. Guo, S. M. Kang, Z. D. Li, T. Tian, J. H. Li, F. D. Wang, P. Yu, S. Y. Chang and Y. Z. Chang, Cell Death Dis., 2023, 14, 161 CrossRef CAS PubMed.
  102. I. Pelizzoni, R. Macco, M. F. Morini, D. Zacchetti, F. Grohovaz and F. Codazzi, Aging Cell, 2011, 10, 172–183 CrossRef CAS PubMed.
  103. J. H. Cheah, S. F. Kim, L. D. Hester, K. W. Clancy, S. E. Patterson, V. Papadopoulos and S. H. Snyder, Neuron, 2006, 51, 431–440 CrossRef CAS PubMed.
  104. NCD Risk Factor Collaboration (NCD-RisC), Lancet, 2021, 398, 957–980 CrossRef PubMed.
  105. P. M. Kearney, M. Whelton, K. Reynolds, P. Muntner, P. K. Whelton and J. He, Lancet, 2005, 365, 217–223 CrossRef PubMed.
  106. Y. L. Liu, K. L. Ai, X. Y. Ji, D. Askhatova, R. Du, L. H. Lu and J. J. Shi, J. Am. Chem. Soc., 2017, 139, 856–862 CrossRef CAS PubMed.
  107. K. K. Griendling, L. L. Camargo, F. J. Rios, R. Alves-Lopes, A. C. Montezano and R. M. Touyz, Circ. Res., 2021, 128, 993–1020 CrossRef CAS PubMed.
  108. M. S. Madhur, F. Elijovich, M. R. Alexander, A. Pitzer, J. Ishimwe, J. P. Van Beusecum, D. M. Patrick, C. D. Smart, T. R. Kleyman, J. Kingery, R. N. Peck, C. L. Laffer and A. Kirabo, Circ. Res., 2021, 128, 908–933 Search PubMed.
  109. Y. Sagara, Y. Hirooka, Y. Kimura, M. Nozoe and K. Sunagawa, Circulation, 2005, 112, U210–U211 Search PubMed.
  110. M. L. Liu, X. Wu, Y. L. Cui, P. L. Liu, B. N. Xiao, X. L. Zhang, J. Zhang, Z. Sun, M. Song, B. Shao and Y. F. Li, Food Chem. Toxicol., 2021, 155, 112388 CrossRef CAS PubMed.
  111. H. L. Li, M. M. Wang, W. L. Kang, Z. M. Lin, F. Gan and K. H. Huang, Toxicology, 2021, 457, 152802 CrossRef CAS PubMed.
  112. J. G. Wang, W. Zhang, Y. Li and L. S. Liu, Nat. Rev. Cardiol., 2023, 20, 531–545 CrossRef PubMed.
  113. A. Gorji, Int. J. Mol. Sci., 2022, 23, 5744 CrossRef PubMed.
  114. D. J. DiSabato, N. Quan and J. P. Godbout, J. Neurochem., 2016, 139, 136–153 CrossRef CAS PubMed.
  115. C. Gao, J. W. Jiang, Y. Y. Tan and S. D. Chen, Signal Transduction Targeted Ther., 2023, 8, 359 CrossRef PubMed.
  116. S. Z. Chen, A. F. U. H. Saeed, Q. Liu, Q. Jiang, H. Z. Xu, G. G. Xiao, L. Rao and Y. H. Duo, Signal Transduction Targeted Ther., 2023, 8, 207 CrossRef PubMed.
  117. J. K. Reid and H. F. Kuipers, Front. Cell. Neurosci., 2021, 15, 704884 Search PubMed.
  118. J. E. Rittiner, M. Moncalvo, O. Chiba-Falek and B. Kantor, Front. Mol. Neurosci., 2020, 13, 148 Search PubMed.
  119. H. S. Kwon and S.-H. Koh, Transl. Neurodegener., 2020, 9, 42 CrossRef PubMed.
  120. R. N. L. Lamptey, B. Chaulagain, R. Trivedi, A. Gothwal, B. Layek and J. Singh, Int. J. Mol. Sci., 2022, 23, 1851 Search PubMed.
  121. Y. K. Li, L. Liu, S. W. Barger and W. S. T. Griffin, J. Neurosci., 2003, 23, 1605–1611 Search PubMed.
  122. Z. H. Liu and H. H. Cheung, Int. J. Mol. Sci., 2020, 21, 8060 CrossRef CAS PubMed.
  123. A. Kanashiro, C. H. Hiroki, D. M. da Fonseca, A. Birbrair, R. G. Ferreira, G. S. Bassi, M. D. Fonseca, R. Kusuda, G. C. M. Cebinelli, K. P. da Silva, C. W. Wanderley, G. B. Menezes, J. C. Alves, A. G. Oliveira, T. M. Cunha, A. S. Pupo, L. Ulloa and F. Q. Cunha, Pharmacol. Res., 2020, 151, 104580 Search PubMed.
  124. Y. H. Guo, H. H. Zeng and C. Gao, Oxid. Med. Cell. Longevity, 2021, 2021, 9931742 CrossRef PubMed.
  125. N. Ochocka and B. Kaminska, Int. J. Mol. Sci., 2021, 22, 3027 CrossRef CAS PubMed.
  126. D. Kempuraj, R. Thangavel, G. P. Selvakumar, S. Zaheer, M. E. Ahmed, S. P. Raikwar, H. Zahoor, D. Saeed, P. A. Natteru, S. Iyer and A. Zaheer, Front. Cell. Neurosci., 2017, 11, 216 CrossRef PubMed.
  127. D. S. Knopman, H. Amieva, R. C. Petersen, G. Chételat, D. M. Holtzman, B. T. Hyman, R. A. Nixon and D. T. Jones, Nat. Rev. Dis. Primers, 2021, 7, 33 CrossRef PubMed.
  128. M. Marković, J. Milošević, W. Wang and Y. Cao, Mol. Pharmacol., 2024, 105, 1–13 CrossRef PubMed.
  129. W. Poewe, K. Seppi, C. M. Tanner, G. M. Halliday, P. Brundin, J. Volkmann, A.-E. Schrag and A. E. Lang, Nat. Rev. Dis. Primers, 2017, 3, 17013 CrossRef PubMed.
  130. L. Chen and M. B. Feany, Nat. Neurosci., 2005, 8, 657–663 Search PubMed.
  131. C. Soto and S. Pritzkow, Nat. Neurosci., 2018, 21, 1332–1340 CrossRef CAS PubMed.
  132. Y. Smith, T. Wichmann, S. A. Factor and M. R. DeLong, Neuropsychopharmacol, 2012, 37, 213–246 CrossRef CAS PubMed.
  133. D. Charvin, R. Medori, R. A. Hauser and O. Rascol, Nat. Rev. Drug Discovery, 2018, 17, 844 CrossRef CAS PubMed.
  134. R. Chen, A. Canales and P. Anikeeva, Nat. Rev. Mater., 2017, 2, 16093 CrossRef CAS PubMed.
  135. E. Pastrana, Nat. Methods, 2010, 8, 24–25 CrossRef.
  136. J. B. Carroll, G. P. Bates, J. Steffan, C. Saft and S. J. Tabrizi, Lancet Neurol., 2015, 14, 1135–1142 CrossRef PubMed.
  137. G. P. Bates, R. Dorsey, J. F. Gusella, M. R. Hayden, C. Kay, B. R. Leavitt, M. Nance, C. A. Ross, R. I. Scahill, R. Wetzel, E. J. Wild and S. J. Tabrizi, Nat. Rev. Dis. Primers, 2015, 1, 15005 Search PubMed.
  138. A. J. Thompson, S. E. Baranzini, J. Geurts, B. Hemmer and O. Ciccarelli, Lancet, 2018, 391, 1622–1636 Search PubMed.
  139. F. Piehl, J. Intern. Med., 2014, 275, 364–381 CrossRef CAS PubMed.
  140. M. Zhou, H. Wang and X. Zeng, Lancet, 2020, 396, 26 CrossRef PubMed.
  141. G. Viegi, S. Maio, S. Fasola and S. Baldacci, J. Aerosol Med. Pulm. Drug Delivery, 2020, 33, 171–177 CrossRef PubMed.
  142. X. Li, X. Cao, M. Guo, M. Xie and X. Liu, BMJ, 2020, 368, m234 CrossRef PubMed.
  143. C. Y. Xu, M. Liu, X. R. Xie, Z. X. Li, Y. F. Zhu, Y. Ye, M. Y. Du, S. H. Hu, T. R. Liu, Y. B. Guo, W. P. Wen, H. L. Liu and Z. X. Tu, Adv. Sci., 2025, 12, e2412626 Search PubMed.
  144. K. Akata and S. F. van Eeden, Int. J. Mol. Sci., 2020, 21, 853 Search PubMed.
  145. R. Shaykhiev, A. Krause, J. Salit, Y. Strulovici-Barel, B. G. Harvey, T. P. O’Connor and R. G. Crystal, J. Immunol., 2009, 183, 2867–2883 CrossRef CAS PubMed.
  146. L. Y. Li, L. Y. Jiao, D. N. Feng, Y. Z. Yuan, X. Q. Yang, J. Li, D. Jiang, H. X. Chen, Q. X. Meng, R. C. Chen, B. X. Fang, X. N. Zou, Z. H. Luo, X. Y. Ye, Y. Hong, C. Liu and C. W. Li, Nat. Commun., 2024, 15, 143 Search PubMed.
  147. F. S. Leitao Filho, H. Takiguchi, K. Akata, S. W. Ra, J.-Y. Moon, H. K. Kim, Y. Cho, K. Yamasaki, S. Milne, J. Yang, C. W. T. Yang, X. Li, C. Nislow, S. F. van Eeden, T. Shaipanich, S. Lam, J. M. Leung and D. D. Sin, Am. J. Respir. Crit. Care Med., 2021, 204, 1143–1152 CrossRef PubMed.
  148. M. Schirmer, A. Garner, H. Vlamakis and R. J. Xavier, Nat. Rev. Microbiol., 2019, 17, 497–511 CrossRef CAS PubMed.
  149. L. Liu, Y. Q. Qin, Z. H. Cai, Y. G. Tian, X. F. Liu, J. S. Li and P. Zhao, Phytomedicine, 2021, 92, 153759 CrossRef CAS PubMed.
  150. C. Brightling and N. Greening, Eur. Respir. J., 2019, 54, 1900651 Search PubMed.
  151. P. J. Murray, Annu. Rev. Physiol., 2017, 79, 541–566 CrossRef CAS PubMed.
  152. A. Papi, C. Brightling, S. E. Pedersen and H. K. Reddel, Lancet, 2018, 391, 783–800 CrossRef PubMed.
  153. M. Sockrider and L. Fussner, Am. J. Respir. Crit. Care Med., 2020, 202, P25–P26 CrossRef PubMed.
  154. X. Yang, Y. Jiang and C. Wang, Mediators Inflammation, 2016, 2016, 1–7 Search PubMed.
  155. M. C. Peters, S. Kerr, E. M. Dunican, P. G. Woodruff, M. L. Fajt, B. D. Levy, E. Israel, B. R. Phillips, D. T. Mauger, S. A. Comhair, S. C. Erzurum, M. W. Johansson, N. N. Jarjour, A. M. Coverstone, M. Castro, A. T. Hastie, E. R. Bleecker, S. E. Wenzel, J. V. Fahy and National Heart, Lung and Blood Institute Severe Asthma Research Program 3, J. Allergy Clin. Immunol., 2019, 143, 104–113 CrossRef CAS PubMed.
  156. C. Michaeloudes, H. Abubakar-Waziri, R. Lakhdar, K. Raby, P. Dixey, I. M. Adcock, S. Mumby, P. K. Bhavsar and K. F. Chung, Mol. Aspects Med., 2022, 85, 101026 CrossRef CAS PubMed.
  157. T. Gui, L. J. Luo, B. Chhay, L. L. Zhong, Y. L. Wei, L. T. Yao, W. Yu, J. Li, C. L. Nelson, A. Tsourkas, L. Qin and Z. L. Cheng, Biomaterials, 2022, 283, 121437 CrossRef CAS PubMed.
  158. P. Lacy, D. A. Latif, M. Steward, S. Musat-Marcu, S. F. P. Man and R. Moqbel, J. Immunol., 2003, 170, 2670–2679 CrossRef CAS PubMed.
  159. X. L. Shi, Z. H. Shi, H. Huang, H. G. Zhu, H. Y. Zhu, D. W. Ju and P. Zhou, Appl. Microbiol. Biotechnol., 2013, 97, 10025–10033 Search PubMed.
  160. M. F. Jabeen, N. D. Sanderson, D. Foster, D. W. Crook, J. L. Cane, C. Borg, C. Connolly, S. Thulborn, I. D. Pavord, P. Klenerman, T. L. Street and T. S. C. Hinks, Microbiol. Spectrum, 2022, 10, e0227921 CrossRef PubMed.
  161. S. B. Hanauer, Inflammatory Bowel Dis., 2006, 12, S3–S9 Search PubMed.
  162. J. A. Keller and K. Hoover, Environ. Entomol., 2023, 52, 759–767 Search PubMed.
  163. J. Cosnes, C. Gower-Rousseau, P. Seksik and A. Cortot, Gastroenterology, 2011, 140, 1785–1794 Search PubMed.
  164. A. L. Richmond, A. Kabi, C. R. Homer, N. Marina-García, K. P. Nickerson, A. I. Nesvizhskii, A. Sreekumar, A. M. Chinnaiyan, G. Nuñez and C. McDonald, Gastroenterology, 2012, 142, 1483–1492.e6 CrossRef CAS PubMed.
  165. M. S. Cairo and A. Beishuizen, Br. J. Haematol.l, 2019, 185, 1021–1042 CrossRef PubMed.
  166. S. R. Vavricka, A. Schoepfer, M. Scharl, P. L. Lakatos, A. Navarini and G. Rogler, Inflammatory Bowel Dis., 2015, 21, 1982–1992 CrossRef PubMed.
  167. M. F. Neurath, Nat. Rev. Immunol., 2014, 14, 329–342 CrossRef CAS PubMed.
  168. D. C. Baumgart and W. J. Sandborn, Lancet, 2012, 380, 1590–1605 Search PubMed.
  169. R. J. Xavier and D. K. Podolsky, Nature, 2007, 448, 427–434 CrossRef CAS PubMed.
  170. J. M. Stormont, A. N. Shah, A. K. Sharma, L. J. Saubermann and R. G. Farmer, Am. J. Gastroenterol., 2013, 108, 1535 Search PubMed.
  171. S. C. Ng, H. Y. Shi, N. Hamidi, F. E. Underwood, W. Tang, E. I. Benchimol, R. Panaccione, S. Ghosh, J. C. Y. Wu, F. K. L. Chan, J. J. Y. Sung and G. G. Kaplan, Lancet, 2017, 390, 2769–2778 CrossRef PubMed.
  172. L. Aldars-García, A. C. Marin, M. Chaparro and J. P. Gisbert, Int. J. Mol. Sci., 2021, 22, 3076 Search PubMed.
  173. J. R. Goldsmith and R. B. Sartor, J. Gastroenterol., 2014, 49, 785–798 CrossRef CAS PubMed.
  174. A. N. Ananthakrishnan, H. Khalili, G. G. Konijeti, L. M. Higuchi, P. de Silva, C. S. Fuchs, J. M. Richter, E. S. Schernhammer and A. T. Chan, Clin. Gastroenterol. Hepatol., 2014, 12, 1879–1886 CrossRef PubMed.
  175. M. F. Neurath, Nat. Rev. Gastroenterol. Hepatol., 2017, 14, 269–278 CrossRef CAS PubMed.
  176. L. Peyrin-Biroulet and M. Lémann, Aliment. Pharmacol. Ther., 2011, 33, 870–879 Search PubMed.
  177. K. Papamichael, A. Gils, P. Rutgeerts, B. G. Levesque, S. Vermeire, W. J. Sandborn and N. Vande Casteele, Inflammatory Bowel Dis., 2015, 21, 182–197 Search PubMed.
  178. L. J. Mohan, J. S. Daly, B. M. Ryan and Z. Ramtoola, Scand. J. Gastroenterol., 2019, 54, 18–26 CrossRef CAS PubMed.
  179. S. Bonovas, G. Fiorino, M. Allocca, T. Lytras, G. K. Nikolopoulos, L. Peyrin-Biroulet and S. Danese, Clin. Gastroenterol. Hepatol., 2016, 14, 1385–1397 CrossRef PubMed.
  180. M. Lemaitre, J. Kirchgesner, A. Rudnichi, F. Carrat, M. Zureik, F. Carbonnel and R. Dray-Spira, JAMA, J. Am. Med. Assoc., 2017, 318, 1679–1686 CrossRef CAS PubMed.
  181. M. Sohail, Mudassir, M. U. Minhas, S. Khan, Z. Hussain, M. de Matas, S. A. Shah, S. Khan, M. Kousar and K. Ullah, Drug Delivery Transl. Res., 2019, 9, 595–614 CrossRef CAS PubMed.
  182. T. Huby and E. L. Gautier, Nat. Rev. Immunol., 2022, 22, 429–443 CrossRef CAS PubMed.
  183. T. Pap and A. Korb-Pap, Nat. Rev. Rheumatol., 2015, 11, 606–615 CrossRef PubMed.
  184. J. S. Smolen and D. Aletaha, Nat. Rev. Rheumatol., 2015, 11, 276–289 CrossRef PubMed.
  185. J. S. Smolen, D. Aletaha, M. Koeller, M. H. Weisman and P. Emery, Lancet, 2007, 370, 1861–1874 CrossRef CAS PubMed.
  186. A. Finckh, B. Gilbert, B. Hodkinson, S. C. Bae, R. Thomas, K. D. Deane, D. Alpizar-Rodriguez and K. Lauper, Nat. Rev. Rheumatol., 2022, 18, 591–602 Search PubMed.
  187. E. M. Gravallese and G. S. Firestein, N. Engl. J. Med., 2023, 388, 529–542 CrossRef CAS PubMed.
  188. N. Nishimoto, K. Ito and N. Takagi, Mod. Rheumatol., 2010, 20, 222–232 CrossRef CAS PubMed.
  189. R. S. Hardy, K. Raza and M. S. Cooper, Nat. Rev. Rheumatol., 2020, 16, 133–144 CrossRef PubMed.
  190. J. M. Hootman, C. G. Helmick, K. E. Barbour, K. A. Theis and M. A. Boring, Arthritis Rheumatol., 2016, 68, 1582–1587 CrossRef PubMed.
  191. R. F. Loeser, S. R. Goldring, C. R. Scanzello and M. B. Goldring, Arthritis Rheumatol., 2012, 64, 1697–1707 CrossRef PubMed.
  192. G. Zhong, X. Yang, X. Jiang, A. Kumar, H. Long, J. Xie, L. Zheng and J. Zhao, Nanoscale, 2019, 11, 23504–23505 Search PubMed.
  193. B. Cillero-Pastor, B. Caramés, M. Lires-Deán, C. Vaamonde-García, F. J. Blanco and M. J. López-Armada, Arthritis Rheumatol., 2008, 58, 2409–2419 Search PubMed.
  194. F. J. Blanco, I. Rego and C. Ruiz-Romero, Nat. Rev. Rheumatol., 2011, 7, 161–169 CrossRef CAS PubMed.
  195. D. B. Burr and A. Utreja, Arthritis Rheumatol., 2018, 70, 157–161 CrossRef PubMed.
  196. H. Fang, L. S. Huang, I. Welch, C. Norley, D. W. Holdsworth, F. Beier and D. Z. Cai, Sci. Rep., 2018, 8, 2855 Search PubMed.
  197. W. H. Hu, Y. Q. Chen, C. Dou and S. W. Dong, Ann. Rheum. Dis., 2021, 80, 413–422 CrossRef CAS PubMed.
  198. M. A. Karsdal, M. Michaelis, C. Ladel, A. S. Siebuhr, A. R. Bihlet, J. R. Andersen, H. Guehring, C. Christiansen, A. C. Bay-Jensen and V. B. Kraus, Osteoarthr. Cartilage, 2016, 24, 2013–2021 CrossRef CAS PubMed.
  199. H. Ujiie, D. Rosmarin, M. P. Schoen, S. Staender, K. Boch, M. Metz, M. Maurer, D. Thaci, E. Schmidt, C. Cole, K. T. Amber, D. Didona, M. Hertl, A. Recke, H. Grasshoff, A. Hackel, A. Schumann, G. Riemekasten, K. Bieber, G. Sprow, J. S. Dan, D. Zillikens, T. Sezin, A. M. Christiano, K. Wolk, R. Sabat, K. Kridin, V. P. Werth and R. J. Ludwig, Front. Med., 2022, 9, 875492 Search PubMed.
  200. J. Schwingen, M. Kaplan and F. C. Kurschus, Int. J. Mol. Sci., 2020, 21, 699 CrossRef CAS PubMed.
  201. T. Bieber, Nat. Rev. Drug Discovery, 2023, 22, 935 CrossRef CAS PubMed.
  202. S. M. Langan, A. D. Irvine and S. Weidinger, Lancet, 2020, 396, 758 Search PubMed.
  203. W. Li, X. Q. Xu, H. Wen, Z. F. Wang, C. Ding, X. C. Liu, Y. X. Gao, H. C. Su, J. X. Zhang, Y. Han, Y. Xia, X. K. Wang, H. Gu and X. Yao, J. Invest. Dermatol., 2019, 139, 1779–1787 Search PubMed.
  204. Y. P. Lai and R. L. Gallo, Trends Immunol., 2009, 30, 131–141 CrossRef CAS PubMed.
  205. C. Roduit, R. Frei, M. Depner, A. M. Karvonen, H. Renz, C. Braun-Fahrländer, E. Schmausser-Hechfellner, J. Pekkanen, J. Riedler, J.-C. Dalphin, E. von Mutius, R. P. Lauener, A. Hyvärinen, P. Kirjavainen, S. Remes, M. Roponen, M.-L. Dalphin, V. Kaulek, M. Ege, J. Genuneit, S. Illi, M. Kabesch, B. Schaub, P. I. Pfefferle and G. Doekes, JAMA Pediatrics, 2017, 171, 655–662 CrossRef PubMed.
  206. K. Piontek, T. Ittermann, A. Arnold, H. Völzke, S. E. Baumeister and C. Apfelbacher, Allergy, 2022, 77, 1915–1917 Search PubMed.
  207. Z. C. Chiesa Fuxench, J. K. Block, M. Boguniewicz, J. Boyle, L. Fonacier, J. M. Gelfand, M. H. Grayson, D. J. Margolis, L. Mitchell, J. I. Silverberg, L. Schwartz, E. L. Simpson and P. Y. Ong, J. Invest. Dermatol., 2019, 139, 583–590 Search PubMed.
  208. P. M. Brunner, J. I. Silverberg, E. Guttman-Yassky, A. S. Paller, K. Kabashima, M. Amagai, T. A. Luger, M. Deleuran, T. Werfel, K. Eyerich, G. Stingl and Councilors of the International Eczema Council, J. Invest. Dermatol., 2017, 137, 18–25 CrossRef CAS PubMed.
  209. R. J. Hay, N. E. Johns, H. C. Williams, I. W. Bolliger, R. P. Dellavalle, D. J. Margolis, R. Marks, L. Naldi, M. A. Weinstock, S. K. Wulf, C. Michaud, C. J. L. Murray and M. Naghavi, J. Invest. Dermatol., 2014, 134, 1527–1534 CrossRef PubMed.
  210. O. Simonetti, T. Bacchetti, G. Ferretti, E. Molinelli, G. Rizzetto, L. Bellachioma and A. Offidani, Antioxidants, 2021, 10, 697 CrossRef CAS PubMed.
  211. Y. B. Jia, J. H. Hu, K. L. An, Q. Zhao, Y. Dang, H. Liu, Z. Wei, S. M. Geng and F. Xu, Nat. Commun., 2023, 14, 2478 Search PubMed.
  212. Z. Liu, Z. Fan, J. X. Liu, J. L. Wang, M. L. Xu, X. L. Li, Y. L. Xu, Y. F. Lu, C. L. Han and Z. H. Zhang, Adv. Sci., 2023, 10, e2204184 CrossRef PubMed.
  213. C. E. M. Griffiths, A. W. Armstrong, J. E. Gudjonsson and J. N. W. N. Barker, Lancet, 2021, 397, 1301–1315 CrossRef CAS PubMed.
  214. K. Ghoreschi, A. Balato, C. Enerbäck and R. Sabat, Lancet, 2021, 397, 754–766 CrossRef CAS PubMed.
  215. S. Subedi, Y. Gong, Y. D. Chen and Y. L. Shi, Drug Des., Dev. Ther., 2019, 13, 2491–2502 Search PubMed.
  216. J. E. Greb, A. M. Goldminz, J. T. Elder, M. G. Lebwohl, D. D. Gladman, J. J. Wu, N. N. Mehta, A. Y. Finlay and A. B. Gottlieb, Nat. Rev. Dis. Primers, 2016, 2, 16082 CrossRef PubMed.
  217. N. Liu, H. Qin, Y. Cai, X. Li, L. Wang, Q. Xu, F. Xue, L. Chen, C. Ding, X. Hu, D. Tieri, E. C. Rouchka, J. Yan and J. Zheng, Ebiomedicine, 2022, 82, 104136 CrossRef CAS PubMed.
  218. Y. X. Sun, X. Liu, L. Wang, L. Xu, K. L. Liu, L. Xu, F. F. Shi, Y. Zhang, N. Gu and F. Xiong, Bioact. Mater., 2022, 10, 117–130 CAS.
  219. R. H. Fiechter, H. M. de Jong, L. J. J. van Mens, I. A. Fluri, S. W. Tas, D. L. P. Baeten, N. G. Yeremenko and M. G. H. van de Sande, Front. Immunol., 2021, 12, 611656 CrossRef CAS PubMed.
  220. S. P. Cannavò, G. Riso, M. Casciaro, E. Di Salvo and S. Gangemi, Free Radical Res., 2019, 53, 829–840 Search PubMed.
  221. C. Conrad, J. Di Domizio, A. Mylonas, C. Belkhodja, O. Demaria, A. A. Navarini, A. K. Lapointe, L. E. French, M. Vernez and M. Gilliet, Nat. Commun., 2018, 9, 25 CrossRef PubMed.
  222. J. Takeshita, S. Grewal, S. M. Langan, N. N. Mehta, A. Ogdie, A. S. Van Voorhers and J. M. Gelfand, J. Am. Acad. Dermatol., 2017, 76, 393–403 CrossRef PubMed.
  223. L. Chen and Z. Shen, Cell. Mol. Immunol., 2020, 17, 64–75 CrossRef CAS PubMed.
  224. L. K. Crawford and M. J. Caterina, Toxicol. Pathol., 2020, 48, 174–189 CrossRef PubMed.
  225. A. Bari, S. Nandyala, J. Balakrishnan, T. Agarwal, T. Dada, R. Saxena and N. Sharma, Indian J. Ophthalmol., 2025, 73, 508–515 Search PubMed.
  226. J. Y. Qi, X. L. Fu, L. Y. Zhang, F. Z. Tan, N. C. Li, Q. H. Sun, X. J. Hu, Z. H. He, M. Xia and R. J. Chai, Fundam. Res., 2025, 5, 192–202 CrossRef CAS PubMed.
  227. S. Pardhan, L. Smith, R. Bourne, A. Davis, N. Leveziel, L. Jacob, A. Koyanagi and G. F. López-Sánchez, Front. Psychol., 2021, 11, 627980 Search PubMed.
  228. Y. L. Wang, F. Cheng, N. N. Hou, Y. T. Tan, S. M. Zhang, Y. B. Hou, W. Guo, J. Peng, W. Li and J. H. Wu, BMC Med., 2025, 23, 118 CrossRef PubMed.
  229. P. Reynard and H. Thai-Van, Therapie, 2024, 79, 283–295 CrossRef PubMed.
  230. S. M. Vlajkovic and P. R. Thorne, Int. J. Mol. Sci., 2021, 22, 5647 CrossRef PubMed.
  231. N. Natarajan, S. Batts and K. M. Stankovic, J. Clin. Med., 2023, 12, 2347 CrossRef CAS PubMed.
  232. K. Fernandez, K. K. Spielbauer, A. Rusheen, L. Z. Wang, T. G. Baker, S. Eyles and L. L. Cunningham, Hear. Res., 2020, 389, 107905 CrossRef CAS PubMed.
  233. V. Ramkumar, D. Mukherjea, A. Dhukhwa and L. P. Rybak, Antioxidants, 2021, 10, 1919 CrossRef CAS PubMed.
  234. W. J. T. Tan and S. M. Vlajkovic, Int. J. Mol. Sci., 2023, 24, 16545 Search PubMed.
  235. S. Li, Z. Y. Lu, Y. Huang, Y. Wang, Q. Jin, X. C. Shentu, J. Ye, J. Ji, K. Yao and H. J. Han, Adv. Sci., 2022, 9, e2200435 CrossRef PubMed.
  236. L. M. Periman, V. L. Perez, D. R. Saban, M. C. Lin and P. Neri, J. Ocul. Pharmacol. Ther., 2020, 36, 137–146 CrossRef CAS PubMed.
  237. S. C. Pflugfelder and C. S. de Paiva, Ophthalmology, 2017, 124, S4–S13 CrossRef PubMed.
  238. S. Seen and L. Tong, Acta Ophthalmol., 2018, 96, E412–E420 CAS.
  239. F. Yu, M. Zheng, A. Y. Zhang and Z. C. Han, J. Controlled Release, 2019, 315, 40–54 CrossRef CAS PubMed.
  240. L. Cui, H. S. Lee, Y. Li, J. H. Choi, J. J. Yun, J. E. Jung, W. Choi and K. C. Yoon, Oxid. Med. Cell. Longevity, 2017, 2017, 4523673 Search PubMed.
  241. F. E. Hakim and A. Farooq, JAMA, J. Am. Med. Assoc., 2022, 327, 478–479 CrossRef PubMed.
  242. G. J. Jaffe, A. D. Dick, A. P. Brézin, Q. D. Nguyen, J. E. Thorne, P. Kestelyn, T. Barisani-Asenbauer, P. Franco, A. Heiligenhaus, D. Scales, D. S. Chu, A. Camez, N. V. Kwatra, A. P. Song, M. Kron, S. Tari and E. B. Suhler, N. Engl. J. Med., 2016, 375, 932–943 Search PubMed.
  243. E. T. E. Cunningham and M. Zierhut, Ocul. Immunol. Inflammation, 2022, 29, 1037–1039 CrossRef PubMed.
  244. L. Ung, U. Pattamatta, N. Carnt, J. L. Wilkinson-Berka, G. Liew and A. J. R. White, Clin. Sci., 2017, 131, 2865–2883 CrossRef CAS PubMed.
  245. J. Kim, J. Chun, M. Ahn, K. Jung, C. Moon and T. Shin, Anat. Cell Biol., 2022, 55, 20–27 Search PubMed.
  246. S. M. Hsu, C. H. Yang, Y. T. Teng, H. Y. Tsai, C. Y. Lin, C. J. Lin, C. C. Shieh and S. H. Chen, Int. J. Mol. Sci., 2020, 21, 3261 CrossRef CAS PubMed.
  247. K. Papangkorn, K. R. Truett, A. T. Vitale, C. Jhaveri, D. K. Scales, C. S. Foster, A. Montieth, J. W. Higuchi, B. Brar and W. I. Higuchi, Curr. Eye Res., 2019, 44, 185–193 CrossRef CAS PubMed.
  248. A. Than, C. H. Liu, H. Chang, P. K. Duong, C. M. G. Cheung, C. J. Xu, X. M. Wang and P. Chen, Nat. Commun., 2018, 9, 4433 CrossRef PubMed.
  249. W. B. Khor, T. Aung, S. M. Saw, T. Y. Wong, P. A. Tambyah, A. L. Tan, R. Beuerman, L. Lim, W. K. Chan, W. J. Heng, J. Lim, R. S. K. Loh, S. B. Lee and D. T. H. Tan, JAMA, J. Am. Med. Assoc., 2006, 295, 2867–2873 CrossRef CAS PubMed.
  250. L. Brown, A. K. Leck, M. Gichangi, M. J. Burton and D. W. Denning, Lancet Infect. Dis., 2021, 21, E49–E57 Search PubMed.
  251. J. Nett, L. Lincoln, K. Marchillo, R. Massey, K. Holoyda, B. Hoff, M. VanHandel and D. Andes, Antimicrob. Agents Chemother., 2007, 51, 510–520 Search PubMed.
  252. J. Berman and D. J. Krysan, Nat. Rev. Microbiol., 2020, 18, 539 CrossRef CAS PubMed.
  253. H. T. Taff, K. F. Mitchell, J. A. Edward and D. R. Andes, Future Microbiol., 2013, 8, 1325–1337 Search PubMed.
  254. M. B. Lohse, M. Gulati, A. D. Johnson and C. J. Nobile, Nat. Rev. Microbiol., 2018, 16, 19–31 CrossRef CAS PubMed.
  255. A. L. Flores-Mireles, J. N. Walker, M. Caparon and S. J. Hultgren, Nat. Rev. Microbiol., 2015, 13, 269–284 CrossRef CAS PubMed.
  256. H. Y. H. Lin, Y. M. Chen, Y. H. Chen, A. P. Ta, H. C. Lee, G. R. MacGregor, N. D. Vaziri and P. H. Wang, Kidney Int., 2021, 99, 870–884 Search PubMed.
  257. S. Tanaka, T. Tanaka and M. Nangaku, Am. J. Physiol. Renal Physiol., 2014, 307, F1187–F1195 Search PubMed.
  258. H. L. Liang, G. Hilton, J. Mortensen, C. P. Johnson and V. Nilakantan, Am. J. Transplant., 2008, 8, 494 Search PubMed.
  259. T. Kalogeris, Y. M. Bao and R. J. Korthuis, Redox Biol., 2014, 2, 702–714 CrossRef CAS PubMed.
  260. H. Kim, C. D. Wrann, M. Jedrychowski, S. Vidoni, Y. Kitase, K. Nagano, C. H. Zhou, J. Chou, V. J. A. Parkman, S. J. Novick, T. S. Strutzenberg, B. D. Pascal, P. T. Le, D. J. Brooks, A. M. Roche, K. K. Gerber, L. Mattheis, W. J. Chen, H. Tu, M. L. Bouxsein, P. R. Griffin, R. Baron, C. J. Rosen, L. F. Bonewald and B. M. Spiegelman, Cell, 2018, 175, 1756–1768 CrossRef CAS PubMed.
  261. R. D. Klein and S. J. Hultgren, Nat. Rev. Microbiol., 2020, 18, 211–226 CrossRef CAS PubMed.
  262. M. Hartmann and T. K. Lindhorst, Eur. J. Org. Chem., 2011, 3583–3609 Search PubMed.
  263. M. Ilkit and A. B. Guzel, Crit. Rev. Microbiol., 2011, 37, 250–261 CrossRef PubMed.
  264. D. W. Denning, M. Kneale, J. D. Sobel and R. Rautemaa-Richardson, Lancet Infect. Dis., 2018, 18, E339–E347 CrossRef PubMed.
  265. Y. Lee, E. Puumala, N. Robbins and L. E. Cowen, Chem. Rev., 2021, 121, 3390–3411 CrossRef CAS PubMed.
  266. J. Berman and P. E. Sudbery, Nat. Rev. Genet., 2002, 3, 918–930 Search PubMed.
  267. M. C. Watson, J. M. Grimshaw, C. M. Bond, J. Mollison and A. Ludbrook, BJOG: Int. J. Obstet. Gynaecol., 2002, 109, 85–95 Search PubMed.
  268. J. N. Krieger, L. Nyberg and J. C. Nickel, JAMA, J. Am. Med. Assoc., 1999, 282, 236–237 CrossRef CAS PubMed.
  269. R. B. Nadler, A. E. Koch, E. A. Calhoun, P. L. Campbell, D. L. Pruden, C. L. Bennett, P. R. Yarnold and A. J. Schaeffer, J. Urol., 2000, 164, 214–218 CrossRef CAS PubMed.
  270. F. F. Pasqualotto, R. K. Sharma, J. M. Potts, D. R. Nelson, A. J. Thomas and A. Agarwal, Urology, 2000, 55, 881–885 CrossRef CAS PubMed.
  271. H. Sung, J. Ferlay, R. L. Siegel, M. Laversanne, I. Soerjomataram, A. Jemal and F. Bray, Ca-Cancer J. Clin., 2021, 71, 209–249 CrossRef PubMed.
  272. S. Falahatkar, E. Shahab, K. G. Moghaddam and E. Kazemnezhad, BJU Int., 2015, 116, 641–649 CrossRef CAS PubMed.
  273. H. S. Jung, P. Verwilst, A. Sharma, J. Shin, J. L. Sessler and J. S. Kim, Chem. Soc. Rev., 2018, 47, 2280–2297 RSC.
  274. L. Li, X. Q. Liang, T. He, X. C. Li, X. Z. Huang, N. Wang, M. L. Shen, Y. Q. Shu, R. Wu, M. M. Zhang, Q. J. Wu and C. Y. Gong, Biomaterials, 2022, 290, 121815 CrossRef CAS PubMed.
  275. J. Li, L. S. Xie, B. Li, C. Yin, G. H. Wang, W. Sang, W. X. Li, H. Tian, Z. Zhang, X. J. Zhang, Q. L. Fan and Y. L. Dai, Adv. Mater., 2021, 33, 2008481 CrossRef CAS PubMed.
  276. Y. Luyu, Z. Jubo and Q. Lunxiu, Eur. J. Immunol., 2019, 49, 578 Search PubMed.
  277. J. Campisi, Annu. Rev. Physiol., 2013, 75, 685–705 Search PubMed.
  278. A. Hernandez-Segura, J. Nehme and M. Demaria, Trends Cell Biol., 2018, 28, 436–453 CrossRef CAS PubMed.
  279. T. Tchkonia, Y. Zhu, J. van Deursen, J. Campisi and J. L. Kirkland, J. Clin. Invest., 2013, 123, 966–972 CrossRef CAS PubMed.
  280. T. Fulop, A. Larbi, G. Dupuis, A. Le Page, E. H. Frost, A. A. Cohen, J. M. Witkowski and C. Franceschi, Front. Immunol., 2018, 8, 1960 CrossRef PubMed.
  281. Y. L. Liu, G. R. Li, T. Z. Zhang, Y. Wu, K. X. Hu, C. Y. Wu, X. Y. Dai, W. Feng and Y. Chen, Adv. Mater., 2025, 37, 2503576 CrossRef PubMed.
  282. X. G. Zhao, Y. X. Yu, X. D. Xu, Z. Q. Zhang, Z. Chen, Y. B. Gao, L. Zhong, J. J. Chen, J. X. Huang, J. Qin, Q. Y. Zhang, X. M. Tang, D. Q. Yang and Z. L. Zhu, Adv. Mater., 2025, 37, 2417536 CrossRef CAS PubMed.
  283. Z. Wei, H. H. Zou, G. Y. Liu, C. H. Song, C. C. Tang, S. Chen, G. R. Zhang, J. C. Ran, Y. F. Wang, X. T. Yin, Y. Cai and W. Han, Bioact. Mater., 2021, 6, 2144–2157 CAS.
  284. X. W. Wang, X. Y. Zhong, Z. B. Zha, G. He, Z. H. Miao, H. L. Lei, Q. Y. Luo, R. Zhang, Z. Liu and L. Cheng, Appl. Mater. Today, 2020, 18, 100464 CrossRef.
  285. J. H. Xie, Q. Q. Li, F. Haesebrouck, L. Van Hoecke and R. E. Vandenbroucke, Trends Biotechnol., 2022, 40, 1173–1194 CrossRef CAS PubMed.
  286. H. F. Li, J. H. Dai, C. Y. Zhao, T. Q. Hu, G. P. Zhao, Q. H. Wang and L. Zhang, Front. Immunol., 2024, 15, 1418717 CrossRef CAS PubMed.
  287. Z. Yi, X. M. Ma, Q. L. Tong, L. Ma, Y. F. Tan, D. N. Liu, C. L. Tan, J. Z. Chen and X. D. Li, Adv. Mater., 2025, 37, 2417534 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.